filename
stringlengths
23
23
markdown
stringlengths
18k
147k
10.1126_science.adf0566
RES EARCH R E S E A R C H A R T I C L E ◥ SLEEP Brain activity of diving seals reveals short sleep cycles at depth Jessica M. Kendall-Bar1,2*, Terrie M. Williams2, Ritika Mukherji3, Daniel A. Lozano2, Julie K. Pitman4, Rachel R. Holser5, Theresa Keates6, Roxanne S. Beltran2, Patrick W. Robinson2, Daniel E. Crocker7, Taiki Adachi2, Oleg I. Lyamin8,9, Alexei L. Vyssotski10, Daniel P. Costa2,5 Sleep is a crucial part of the daily activity patterns of mammals. However, in marine species that spend months or entire lifetimes at sea, the location, timing, and duration of sleep may be constrained. To understand how marine mammals satisfy their daily sleep requirements while at sea, we monitored electroencephalographic activity in wild northern elephant seals (Mirounga angustirostris) diving in Monterey Bay, California. Brain-wave patterns showed that seals took short (less than 20 minutes) naps while diving (maximum depth 377 meters; 104 sleeping dives). Linking these patterns to accelerometry and the time-depth profiles of 334 free-ranging seals (514,406 sleeping dives) revealed a North Pacific sleepscape in which seals averaged only 2 hours of sleep per day for 7 months, rivaling the record for the least sleep among all mammals, which is currently held by the African elephant (about 2 hours per day). A cross the animal kingdom, sleep is crit- ical for energy conservation, immune function, memory, and learning (1). Dis- ruptions to sleep, including obstructive sleep apnea and shift work, negatively affect human health (2, 3). By comparison, diverse sleeping habits among wild animals reflect adaptations to resolve conflicts between sleeping or feeding while avoiding predation and exhaustion (4–6). In response to these trade-offs, cows sleep-chew, horses sleep- stand, ostriches sleep-stare, and frigate birds sleep-fly (7–10). Marine mammals face unique challenges in obtaining adequate daily sleep (11). Most of them feed underwater and breathe at the ocean surface, where predators typically attack (12). Activity budgets of mammals at sea reflect the balance between these survival needs, which often push the animals toward physiological extremes such as large body size, prolonged activity, and enhanced oxygen stores (13). For example, northern elephant seals (Mirounga angustirostris) travel >10,000 km during 7-month-long foraging trips. Seals minimize time at the surface (~2 min between 10- to 30-min dives) to reduce predation risk by killer 1Scripps Institution of Oceanography, University of California San Diego, San Diego, CA, USA. 2Ecology and Evolutionary Biology, University of California Santa Cruz, Santa Cruz, CA, USA. 3Department of Neuroscience, University of Oxford, Oxford, UK. 4Sleep Health MD, Santa Cruz, CA, USA. 5Institute of Marine Sciences, University of California Santa Cruz, Santa Cruz, CA, USA. 6Ocean Sciences Department, University of California Santa Cruz, Santa Cruz, CA, USA. 7Department of Biology, Sonoma State University, Rohnert Park, CA, USA. 8Semel Institute for Neuroscience and Human Behavior, University of California Los Angeles, Los Angeles, CA, USA. 9A.N. Severtsov Institute of Ecology and Evolution, Moscow, Russia. 10Institute of Neuroinformatics, University of Zurich and Swiss Federal Institute of Technology (ETH), Zurich, Switzerland. *Corresponding author. Email: jkb@ucsc.edu whales and white sharks while maximizing foraging time (12, 14–16). They also feed around the clock on small prey to satisfy the energy requirements associated with their large body size (17). Given these ecophysiological demands, a long-standing question has been when, where, and how do seals sleep at sea? A new tool to detect sleep at sea We developed a new submersible system to record brain activity (electroencephalogram, EEG) and heart rate (electrocardiogram, ECG) concurrently with dive depth and motion of elephant seals at sea [Fig. 1E; (18)]. These sen- sors identified sleep states [rapid eye move- ment (REM) sleep and slow-wave sleep (SWS); figs. S1 to S3 and table S2], swimming effort (stroke rate), and three-dimensional (3D) diving behavior in freely moving female ju- venile seals (n = 13 seals) (18). We recorded sleep in a controlled laboratory environment (n = 5 seals) and in the wild (n = 8 seals) at four locations, including on the beach, in shal- low water, offshore along the continental shelf (depth <250 m), and in the open ocean (depth >250 m; table S1). EEG recordings allowed us to pair sleep states with diving behavior recorded in time-depth profiles for juveniles over multiple days at sea (104 sleeping dives). We used these sleep signatures to estimate sleep patterns across >3 million dives by 334 free-ranging adult females over prolonged trips at sea (514,406 sleeping dives across 53,581 recording days). On a typical sleeping dive, seals transitioned from an awake glide into SWS. Although asleep, they could maintain their upright posture for several minutes (Fig. 1 and movie S1). These results underscore the importance of EEG in assessing sleep state (18). As seals shifted from SWS to REM sleep, sleep paraly- sis resulted in a loss of postural control. Seals turned upside down and drifted downwards in a “sleep spiral.” Sleep spirals tightened from a median diameter of 7.5 ± 7.9 m [median ± interquartile range (IQR)] at 71 ± 97 s (median ± IQR) in SWS to 3.3 ± 3.5 m loops at 40 ± 29 s in REM (Fig. 1). Sleep spirals consisted of two to 13 consecutive 360-degree loops at 82 to 377 m depth. On the continental shelf, seals slept motionless on the ocean floor at 64 to 249 m depth. The predation risk of sleep at sea Among marine mammals, unihemispheric sleep (SWS in only one hemisphere) allows captive cetaceans and otariids (fur seals and sea lions) to swim and keep one eye open during sleep (11, 19). This suggests that cetaceans and otariids can sleep while monitoring predators (20, 21). Unihemispheric sleep has not been detected in captive true seals (family Phocidae) such as elephant seals (22). Similarly, our study did not reveal sleep asymmetry between hemi- spheres (<2-fold difference). This suggests that true seals use an alternative solution to mit- igate predation risk. This study experimen- tally confirms the hypothesis (22, 23) that in the absence of unihemispheric sleep, elephant seals’ extreme diving abilities allow sleep deep below the ocean surface, out of view of visual predators. The sleep paralysis that co-occurs with REM sleep would make seals especially vulnerable to predation (1). REM is often minimized for aquatic mammals because the accompanying paralysis can also prevent access to air (22). In captive fur seals confined to water, REM is virtually eliminated (24). Elephant seals at sea reduce REM sleep, as is seen in captive fur seals and true seals in water (24–27), but unexpec- tedly exhibit a large proportion of REM [26.5 ± 5.0% (mean ± SD) in total sleep time over- all and 29.1 ± 4.3% of at-sea total sleep time; table S3]. This compares to 11%, 6%, 5%, and 1% in aquatic sleep for captive Caspian seals, harp seals, walruses, and fur seals, respectively [(22, 24–27); see the materials and methods). Our at-sea deployments occurred during late spring, when juvenile seal aggregations at- tract predators (14). While transiting over the continental shelf, juvenile seals alter their swimming behavior to avoid predation (28). Unexpectedly, we found that seals slept pro- portionally more on the continental shelf than in the open ocean (Figs. 2 and 3A). One seal performed up to 36 consecutive sleeping dives on the continental shelf but fewer than five at sea. This suggests that seals can safely sleep at depth despite elevated coastal predation risk. Finding time to sleep at sea Without unihemispheric sleep allowing con- tinuous vigilance, seals are vulnerable and Kendall-Bar et al., Science 380, 260–265 (2023) 21 April 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. 3D drift dive sleep patterns. (A) A 23-min sleeping dive showing stroke rate [strokes per minute (spm)], heart rate [beats per minute (bpm)], left (L) EEG (mV), right (R) EEG (mV), L EEG spectrogram [power (dB) for frequency (Hz) over time], pitch (radians), roll (absolute value; radians) and heading (radians), and time (minutes of dive). (B to D) Raw EEG and ECG signals during the transition to light SWS (B), deep SWS (C), and REM (D). During SWS, high-voltage, low-frequency slow waves are present. During REM, low-voltage, high-frequency EEG activity co-occurs with heart rate variability. (E) EEG logger configuration demonstrating headcap and logger placement. (F) Schematic demonstrating placement of electrodes for electrooculogram (EOG), EEG, ECG, and electromyogram (EMG). (G) 3D dive profile color-coded by sleep state: Active waking is shown in dark blue, quiet waking in light blue, light SWS in light green, deep SWS in teal, and REM in yellow. (H) Top view of sleep spiral. (I) Depth over time showing nested durations of gliding, electrophysiological sleep, constant drift rate, and spiraling. Kendall-Bar et al., Science 380, 260–265 (2023) 21 April 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Sleep patterns from land to sea. (A) Daily sleep quotas for seals in the laboratory (on land and in the pool) and in the wild (on land, in shallow water, on the continental shelf, and in the open ocean), including active waking (dark blue), calm (lighter blue), drowsiness (purple), REM sleep (yellow), and SWS (light blue). REM sleep totals include certain and putative REM (see “REM scoring” section in the supplementary materials). (B) Schematic showing the resting postures of seals in each habitat, including seals resting on the ocean floor on the continental shelf and drifting in the open ocean. (C) 2D map with bathymetry showing georeferenced dead-reckoned tracks for three animals recorded at sea. (D) 3D map demonstrating sleeping dive sequence, including the sleeping dive from Fig. 1. Kendall-Bar et al., Science 380, 260–265 (2023) 21 April 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Sleep identification model performance. Time-depth records for two juvenile seals (A and B) are colored to indicate surface intervals (light blue), dives (dark blue), glides (blue), SWS (green), and REM sleep (yellow). In panels (A1) and (B1), the identified sleep segments are denoted below the dive profile, where outlined dots at the beginning and end of sleep segments are colored from yellow to dark blue according to overlap with sleep (“percent nap overlap”). Light shaded regions above the dive profile in panels (A1) and (B1) demonstrate sleep identification accuracy (false positives in blue, false negatives in yellow, and true positives in green). Panels (A2) and (B2) display EEG spectrograms and heart rate for two adjacent sleeping dives. Panels (A3) and (B3) quantify daily activity budgets (or provide estimates) in hours per day of diving, sleep estimates (upper bound – unfiltered sleep ID; best estimate/ lower bound includes only sleep ID segments that meet filter criteria), gliding (long glides >200 s), sleeping (both SWS and REM), and REM sleep. Kendall-Bar et al., Science 380, 260–265 (2023) 21 April 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Estimating daily sleep for 334 adult females. (A) Map showing estimated sleep (hours per day) for two of 15 seals instrumented with stroke rate loggers (basemap matches bathymetry legend in Fig. 2). Each circle represents 1 day, with circle size and color reflecting daily sleep (hours per day). (B1 and B2) Time- depth records for two 24-hour days far from shore (B1) and close to Vancouver Island (B2), demonstrating the difference in sleep during pelagic versus coastal foraging. Yellow dots and green shaded regions above dive profiles indicate identified sleep segments that 100% overlap with a long glide segment (yellow and blue shading represent false negatives and false positives, respectively). (C) Map with spatially averaged daily sleep estimates clipped to the extent of tracking data (1 point per seal day) across 342 good-quality tracks from 267 adult females. Note the higher sleep time along the coast and foraging grounds. (D) Daily activity budgets for long postmolt foraging trips (n = 164 seals) including surface intervals, diving, long glides, long drifts, sleep estimates (filtered long drifts), and long surface intervals. Sleep estimates demonstrate low sleep time throughout the trip. (E) Comparative figure showing total sleep time in terrestrial mammalian carnivores, omnivores, and herbivores [reprinted with permission (1)]. Extremes of sleep time on land and at sea from EEG recordings in juveniles (Fig. 2) are plotted for comparison. Sleep durations for other mammals are based on behavior and/or EEG in the laboratory and/or wild. Differences in recording location (laboratory versus wild) and sleep identification technique (EEG versus behavior) complicate sleep quantification and direct comparison. unable to actively transit during sleep to max- imize foraging efficiency. Between heightened predation risk and lost foraging opportunities, we expected sleep to be strongly restricted at sea. Supporting this hypothesis, we discovered that seals’ daily sleep time was >5 times higher on land than at sea (Fig. 2). Seals slept up to 14 hours/day on land [10.8 ± 3.0 hours/day (mean ± SD)] but as little as 0 hours/day at sea (1.7 ± 0.7 hours/day; tables S3 and S4). After returning from 2 to 3 days at sea, seals remained on land for 18 to 43 hours, sleeping up to 53.3% of each hour before returning to shallow water (fig. S4). This moderate sleep rebound was comparable to the daily patterns Kendall-Bar et al., Science 380, 260–265 (2023) 21 April 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E of other seals at the colony (fig. S4), suggesting that this relatively short multiday trip did not incur a notable sleep debt. Substantial fluctuations in sleep duration allow birds to prioritize migration and breed- ing for several days (5, 10). As mammals, ele- phant seals similarly partition strategies over long time scales, sacrificing sleep at sea to sup- port the energy requirements associated with their large size and deferring sleep until they leave the water and are on land with no preda- tors. Seals in laboratory settings show modest differences between total sleep time on land and water (24, 25). Here, we demonstrate greater ex- tremes in total sleep time for wild seals that are necessary to balance sleep with the need to replenish the energy stores of a large, highly mobile predator at sea. Mapping range-wide sleep patterns at the population level Using the paired electrophysiological and time-depth signatures of SWS and REM sleep from instrumented seals (Figs. 1 and 2), we developed a high-accuracy sleep identifica- tion algorithm that identified segments of inactivity characterized by low vertical speed and acceleration from time-depth data (93% accuracy; Fig. 3 and figs. S5 to S7). This algo- rithm allowed us to estimate sleep quotas for 334 adult seals from their diving data recorded over several months at sea [n = 170 short trips (74.6 ± 9.5 days) and n = 164 long trips (217.7 ± 24.7 days)] (Fig. 4). These analyses indicated that daily sleep quotas were likely to be uni- versally low (1.1 ± 1.1 hours/day for short 70-day trips and 2.2 ± 1.6 hours/day for >200-day trips) (Fig. 4). Expanding this analysis to the population level, we can map range-wide sleep patterns to identify critical habitats for protecting wild seals while they sleep at sea. These “sleep- scapes,” which are based on 342 foraging trips by seals across the North Pacific (Fig. 4C), re- veal the same unexpected sleep patterns as in juvenile EEG records. That is, seals slept more while closer to the coastline despite greater predation risk (Fig. 4B1) (14). Because coastal foragers consume fewer, larger prey (17), our findings suggest that these seals must either expend more energy hunting for larger prey or require more time to rest and process such prey. Although the coastal water column may harbor more predators, the continental shelf may also facilitate sleep by providing shelter from predators and relative proximity to the surface. These findings and the resulting sleep- scape aid in identifying critical habitats that may guide coastal conservation efforts for wild animals. By connecting locomotion with different forms of sleep (SWS versus REM) in northern elephant seals, the present study provides conclusive evidence of sleep during drift dives (23, 29, 30). Furthermore, these unique record- ings of brain activity for a wild, free-ranging marine mammal at sea show that sleeping at depth allows seals to drift safely in and out of sleep paralysis. However, these respites are short, because the large body size [456 to 687 kg (min-max adult female arrival breeding mass); (31)] of this elite diver that forages and sleeps in the dark must be sustained by near-constant foraging at sea. Sleep patterns interpreted from the dive records of hundreds of seals re- vealed only 2 hours/day of sleep for months, rivaling the record for the least sleep among mammals [the African elephant at 2 hours per day; (32)]. Both this method (applying sleep signatures from a small sample to reveal population-level patterns) and these find- ings (a detailed understanding of sleep for a highly mobile, large mammal) provide oppor- tunities for understanding sleep’s function, evolution, and pathology across mammals, including in humans. RE FERENCES AND NOTES J. M. Siegel, Nature 437, 1264–1271 (2005). 1. 2. C. V. Senaratna et al., Sleep Med. Rev. 34, 70–81 (2017). 3. C. M. Almeida, A. Malheiro, Sleep Sci. 9, 164–168 (2016). 4. C. A. Wyse, A. N. Coogan, C. Selman, D. G. Hazlerigg, J. R. Speakman, Biol. Lett. 6, 696–698 (2010). J. A. Lesku et al., Science 337, 1654–1658 (2012). 5. 6. N. C. Rattenborg, S. L. Lima, C. J. Amlaner, Nature 397, 397–398 (1999). 7. E. Ternman, L. Hänninen, M. Pastell, S. Agenäs, P. P. Nielsen, Appl. Anim. Behav. Sci. 140, 25–32 (2012). 8. T. Belling, Equine Practice 12, 2–26 (1990). 9. J. A. Lesku et al., PLOS ONE 6, e23203 (2011). 10. N. C. Rattenborg et al., Nat. Commun. 7, 12468 (2016). 11. O. I. Lyamin, P. R. Manger, S. H. Ridgway, L. M. Mukhametov, J. M. Siegel, Neurosci. Biobehav. Rev. 32, 1451–1484 (2008). 12. R. A. Martin, N. Hammerschlag, Mar. Biol. Res. 8, 90–94 (2012). 13. M. J. Weise, D. P. Costa, J. Exp. Biol. 210, 278–289 (2007). 14. S. J. Jorgensen et al., Sci. Rep. 9, 6153 (2019). 15. P. W. Robinson et al., PLOS ONE 7, e36728 (2012). 16. N. Hammerschlag, R. A. Martin, C. Fallows, Environ. Biol. Fishes 76, 341–350 (2006). 17. T. Adachi et al., Sci. Adv. 7, eabg3628 (2021). 18. J. M. Kendall-Bar et al., Anim. Biotelem. 10, 16 (2022). 19. N. C. Rattenborg, C. J. Amlaner, S. L. Lima, Neurosci. Biobehav. Rev. 24, 817–842 (2000). 20. J. M. Kendall-Bar, A. L. Vyssotski, L. M. Mukhametov, J. M. Siegel, O. I. Lyamin, PLOS ONE 14, e0217025 (2019). 21. O. I. Lyamin et al., Behav. Brain Res. 129, 125–129 (2002). 22. O. I. Lyamin, J. M. Siegel, Handb. Behav. Neurosci. 30, 375–393 (2019). 23. Y. Mitani et al., Biol. Lett. 6, 163–166 (2010). 24. O. I. Lyamin et al., Curr. Biol. 28, 2000–2005.e2 (2018). 25. O. I. Lyamin, J. Sleep Res. 2, 170–174 (1993). 26. O. I. Lyamin et al., Dokl. Biol. Sci. 444, 188–191 (2012). 28. B. J. Le Boeuf, D. E. Crocker, in Great White Sharks, The Biology of Carcharodon carcharias, A. P. Klimley and D. G. Ainley, Eds. (Academic Press, 1996), pp. 193–205. 29. D. E. Crocker, B. J. L. Boeuf, D. P. Costa, Can. J. Zool. 75, 27–39 (1997). 30. Y. Naito et al., Funct. Ecol. 27, 710–717 (2013). 31. R. R. Holser, D. E. Crocker, P. W. Robinson, R. Condit, D. P. Costa, Proc. Biol. Sci. 288, 20211258 (2021). 32. N. Gravett et al., PLOS ONE 12, e0171903 (2017). 33. Data for: J. M. Kendall-Bar et al., Brain activity of diving seals reveals short sleep cycles at depth, Dryad (2023); https://doi.org/10.7291/D1ZT2B. 34. Code for: J. M. Kendall-Bar et al., Brain activity of diving seals reveals short sleep cycles at depth, Zenodo (2023); https://doi.org/10.5281/zenodo.7702650. AC KNOWLED GME NTS We acknowledge the students, volunteers, and researchers contributing to the long-term Año Nuevo elephant seal research program, especially C. Kuhn, J. Hassrick, S. Simmons, M. Fowler, S. Peterson, A. Favilla, S. Kienle, and L. Hückstädt for their assistance in collecting female tracking data. We also thank volunteers C. Lopez and J. Nichols for their help with data collection and analysis; Año Nuevo State Park and the Año Nuevo UC Natural Reserve for their ongoing support, TMMC veterinarians C. Field and S. Johnson for assistance with pilot studies; and Long Marine Lab staff and volunteers for facilitating lab-based studies. Engineers and technicians P. Guerrero, E. Slattery, J. Bielke, and colleagues at Ocean Innovations, Scripps Institution of Oceanography, and the Shorter laboratory at the University of Michigan assisted with the development of the tag housing. Funding: This work was supported by the National Ocean Partnership Program (grant N00014-02-1-1012 to D.P.C.); the National Science Foundation (grant N1656282 to D.P.C.); the Strategic Environmental Research and Development Program (SERDP grant RC20-C2-1284 to D.P.C.); the Office of Naval Research (grants N00014-18-1-2822, N00014-00-1-0880, N00014-03-1-0651, and N00014-08-1-1195 to D.P.C. and grant N00014-20-1-2762 to T.M.W.); the Office of Naval Research Defense University Research Instrumentation Program (grant N00014-19-1-2178 to T.M.W. and J.M.K.-B.); the E&P Sound and Marine Life Joint Industry Project (JIP) of the International Association of Oil and Gas Producers (IOGP grant JIP2207-23 to D.P.C.); a National Geographic Early Career Grant (J.M.K.-B.); a Steve & Rebecca Sooy Graduate Research Fellowship (J.M.K.-B.); Achievement Rewards for College Scientists (J.M.K.-B.); the National Science Foundation Graduate Research Fellowship Program (J.M.K.-B.); and a Special Research Grant from the Committee on Research at UC Santa Cruz (D.P.C. and J.M.K.-B.). Author contributions: Conceptualization: J.M.K.-B., T.M.W., D.P.C.; Funding acquisition: J.M.K.-B., T.M.W., D.P.C.; Investigation: J.M.K.-B., R.M., D.A.L., J.K.P., R.R.H., T.K., R.S.B., P.W.R., D.E.C., T.A.; Methodology: J.M.K.-B., A.L.V.; Supervision: T.M.W., D.P.C.; Visualization: J.M.K.-B.; Writing – original draft: J.M.K.-B., T.M.W., D.P.C.; Writing – review and editing: J.M.K.-B., T.M.W., R.M., D.A.L., J.K.P., R.R.H., T.K., R.S.B., P.W.R., D.E.C., T.A., O.I.L., A.L.V., D.P.C. Competing interests: The authors declare no competing interests. Data and materials availability: Statistical data are presented primarily in the main text and online supplementary materials, with electronic versions of original sleep data and analysis scripts directly available online through Dryad (33) and Zenodo (34), respectively. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf0566 Materials and Methods Figs. S1 to S7 Tables S1 to S5 References (35–60) Movie S1 MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. 27. O. I. Lyamin, P. O. Kosenko, J. L. Lapierre, L. M. Mukhametov, J. M. Siegel, J. Neurosci. 28, 12614–12621 (2008). Submitted 30 September 2022; accepted 14 March 2023 10.1126/science.adf0566 Kendall-Bar et al., Science 380, 260–265 (2023) 21 April 2023 6 of 6
10.1126_science.adf2698
RES EARCH SEMICONDUCTORS Room-temperature wavelike exciton transport in a van der Waals superatomic semiconductor Jakhangirkhodja A. Tulyagankhodjaev, Petra Shih†, Jessica Yu†, Jake C. Russell, Daniel G. Chica, Michelle E. Reynoso, Haowen Su, Athena C. Stenor, Xavier Roy, Timothy C. Berkelbach, Milan Delor* The transport of energy and information in semiconductors is limited by scattering between electronic carriers and lattice phonons, resulting in diffusive and lossy transport that curtails all semiconductor technologies. Using Re6Se8Cl2, a van der Waals (vdW) superatomic semiconductor, we demonstrate the formation of acoustic exciton-polarons, an electronic quasiparticle shielded from phonon scattering. We directly imaged polaron transport in Re6Se8Cl2 at room temperature, revealing quasi-ballistic, wavelike propagation sustained for a nanosecond and several micrometers. Shielded polaron transport leads to electronic energy propagation lengths orders of magnitude greater than in other vdW semiconductors, exceeding even silicon over a nanosecond. We propose that, counterintuitively, quasi- flat electronic bands and strong exciton–acoustic phonon coupling are together responsible for the transport properties of Re6Se8Cl2, establishing a path to ballistic room-temperature semiconductors. S emiconductor technologies rely on trans- porting energy and information carriers, often in the form of electrons or excitons (bound electron-hole pairs), from source to target. At room temperature, these carriers rapidly scatter with lattice vibrations (phonons) on nanometer and femtosecond scales. Scattering leads to electronic energy dissipation, joule heating, and loss of phase co- herence and directionality, imposing strict speed and efficiency limits on all semiconductor tech- nologies. Breaking through these limits requires semiconductors that sustain ballistic (scatter- free), wavelike flow of energy over macroscopic distances at room temperature, a long-sought goal that would enable ballistic transistors (1), low-loss energy harvesting, and wave-based information technologies (2). Here, we demonstrate macroscopic, wavelike exciton flow at room temperature in the van der Waals (vdW) superatomic material Re6Se8Cl2 (Fig. 1A). Re6Se8Cl2 is a semiconductor with an indirect bandgap of 1.6 eV and an exciton binding energy of ~100 meV (3). Its superatom building blocks consist of Re6 octahedra en- closed in Se8 cubes. Each Re6Se8 unit is co- valently bonded to four neighbors to form a two-dimensional (2D) pseudosquare lattice capped by Cl atoms at the apical positions. The Re6Se8Cl2 layers stack out of plane to create a bulk vdW crystal with weak interlayer electron- ic coupling (4). The crystal can be exfoliated to the monolayer limit, which is advantageous for integration in gated devices (3, 5, 6). Re6Se8Cl2 exhibits relatively weak intercluster electronic coupling, as evidenced by the elec- tronic band structure (Fig. 1B) (3, 7). Strong coupling of electrons to intercluster optical Department of Chemistry, Columbia University, New York, NY 10027, USA. *Corresponding author. Email: milan.delor@columbia.edu †These authors contributed equally to this work. phonons (8) leads to further band flattening at room temperature (7) and has been impli- cated in the emergence of superconductivity in this material and related classes (6, 9). In this work, we demonstrate that these quasi-flat electronic bands, in combination with strong coupling to acoustic phonons, lead to the for- mation of acoustic exciton-polarons (Fig. 1C), quasiparticles of excitons bound to an acoustic lattice deformation. Through direct imaging of polaron propagation, we reveal that they are shielded from lattice scattering, leading to quasi- ballistic transport over several micrometers at room temperature, currently limited only by crystal size. Our observations challenge the common notion that strong electronic cou- pling is required for long-range transport. Exciton transport in Re6Se8Cl2 is quasi-ballistic We directly imaged exciton transport in single- crystal Re6Se8Cl2 using ultrafast stroboscopic scattering microscopy (stroboSCAT) (10–12) (Fig. 1D and figs. S1 to S3). An above-gap, diffraction-limited visible pump generates excitons, and then a backscattering widefield probe (1.55 eV) slightly below the electronic bandgap spatially resolves how the excitons modify the local polarizability of the material. By varying the pump-probe time delay, we spatiotemporally tracked the evolution of photo- excitations in an all-optical, noninvasive, and contact-free measurement. Figure 1E and movie S1 display representative stroboSCAT data obtained in a 60-nm-thick Re6Se8Cl2 flake pre- pared by mechanical exfoliation (13). Two key features emerge from these data: First, the initial negative (dark) stroboSCAT contrast turns to positive (bright) contrast on picosecond timescales, which, as discussed below (Fig. 2), represents a transition from a bare exciton to an exciton-polaron. Second, the exciton-polaron propagates several micrometers to the edge of the flake in less than a nanosecond. This very fast and long-range transport differs starkly from exciton transport in other molecular or 2D semiconductors (table S1). For comparison, Fig. 1F displays stroboSCAT data of exciton trans- port in the archetypal vdW semiconductor WSe2 (bilayer flake on glass; see fig. S4 for monolayer and bulk data), exhibiting much slower and shorter-range transport. These results are coun- terintuitive, given that the effective mass of ex- citons in WSe2 is much smaller than in Re6Se8Cl2 (table S1). To quantify and rationalize the exceptional transport properties of Re6Se8Cl2, we plotted the mean squared displacement (MSD) of the photoexcited population profile observed in stroboSCAT as MSD = s2(t) − s2(0), where s is the Gaussian width of the population den- sity profile at time delay t (13). Figure 1G compares the MSD for exciton-polarons in Re6Se8Cl2 against the MSD for excitons in bilayer WSe2 and charge carriers in intrinsic monocrystalline Si (14), which exhibit some of the best transport among 2D and 3D semi- conductors, respectively. We find that the MSD at 1.1 ns in Re6Se8Cl2 is 23 times that in bulk WSe2, 65 times that in bilayer WSe2, 120 times that in monolayer WSe2, and almost twice that of electrons in intrinsic Si and the recently reported cubic boron arsenide (15, 16) (table S1). On the basis of the 11-ns polaron lifetime (fig. S5), we estimate that the polaron propa- gation length in Re6Se8Cl2 would exceed 25 mm in the absence of crystal boundaries. The superlinear behavior of the MSD for Re6Se8Cl2 differs from the linear behavior in Si and WSe2. We fit the MSD to a power law, MSD º ta (11, 12, 17). In the limit of diffusive transport, where scattering lengths are much shorter than the propagation length, a ¼ 1. This regime, exemplified by the linear MSD for Si and WSe2 in Fig. 1F, is observed in virtually all semiconductors beyond the first few femto- seconds after photoexcitation (18). In the limit of coherent, ballistic transport (no scattering), a = 2, meaning that distance is proportional to time with a slope that defines the velocity. In Re6Se8Cl2, we observe quasi-ballistic transport (a = 1.67 ± 0.13) (Fig. 1G) sustained for nano- seconds, until the flake edge is reached. Plot- ting the same data as distance versus time provides an effective propagation velocity of 2.3 km/s (fig. S6). Monte Carlo simulations re- producing the observed MSD yield an exciton– lattice scattering time of 215 ps, indicating an extraordinary mean free path between scattering events exceeding 1 mm for exciton-polarons in Re6Se8Cl2 (Fig. 1H; see figs. S7 and S8 for simulation details and alternative models). These findings reveal that the mechanism for fast and long-range transport in Re6Se8Cl2 is effi- cient shielding from scattering, amply compen- sating for the large effective mass (and thus low intrinsic velocity) of the polaron. These results are reproducible in multiple flakes of different Tulyagankhodjaev et al., Science 382, 438–442 (2023) 27 October 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Imaging exciton transport in Re6Se8Cl2. (A) Crystal structure of Re6Se8Cl2. (B) Band structure of Re6Se8Cl2 calculated using density functional theory with the Perdew-Burke-Ernzerhof (PBE) exchange correlation functional (13). (C) Formation of acoustic polarons through a deformation potential interaction. (D) Schematic for optical far-field imaging of polaron transport (details in the text and figs. S1 to S3). (E) Picosecond stroboSCAT time series displaying exciton (dark contrast) and exciton-polaron (bright contrast) propagation in a 60-nm-thick Re6Se8Cl2 flake on glass. DI/I refers to the pump-induced change in backscattered light intensity. (F) Exciton propagation in bilayer WSe2 on glass. (G) Mean squared displacement (MSD) of exciton-polarons in Re6Se8Cl2 (red), excitons in WSe2 (blue), and charge carriers in Si (gray and black). Error bars are 1 SD. Only Re6Se8Cl2 displays superlinear behavior, indicating superdiffusive transport characterized by the exponent a. The box plot shows the spread of a values across 11 different datasets, indicating mean and median values of 1.67 and 1.7, respectively. (H) Monte Carlo simulation of a for different particle velocities (v) and scattering times (t) for our experimental configuration. The circle corresponds to simulation parameters of v = 5.5 km/s and t = 215 ps that reproduce the experimental MSD. The black contour traces a = 1.67. (I) Illustration of quasi-ballistic motion of polarons in Re6Se8Cl2 compared with diffusive motion for excitons in WSe2 and other semiconductors. Data acquired at 295 K. thicknesses (Figs. 1 and 3), for both above-gap and band-edge pump excitation (fig. S9), for pump temporal pulse widths spanning 60 fs to 60 ps (fig. S10), and across a range of fluences explored in detail below (Fig. 3), indicating that neither hot carrier transport (19), phonon winds (20, 21), nonlinear recombination (22), thermal gradients (23), nor strain waves (24–27) are responsible for the observed behavior (figs. S9 to S12 and table S2). Excitons in Re6Se8Cl2 form exciton-polarons To confirm that polaron formation is respon- sible for the observed transport behavior, we tracked the energetic evolution of excitons after photoexcitation and correlated these dynamics to optical transport measurements. Figure 2A displays transient reflectance spectra in the re- gion of the semiconductor band edge, exhibiting a bleach around 1.57 eV and a photoinduced absorption ~90 meV higher in energy. The primary dynamic evolution observed is a spec- tral redshift on a 1.5-ps timescale that convolves all peaks. Identical dynamics are observed for near-band-edge excitation (fig. S9), ruling out the possibility that electronic thermalization is responsible for the redshift. Tracking the zero- crossing of the transient reflectance profile (dashed line in Fig. 2A) provides a handle on redshift kinetics, plotted in Fig. 2B. We observe an overall 48 meV redshift with an evolution re- sembling a strongly damped oscillator (red line in Fig. 2B). These dynamics echo those previous- ly observed for (large) polaron formation (28–31), wherein energetic stabilization occurs over a sin- gle vibrational period of the associated lattice de- formation. This spectral evolution is responsible for the switch from dark to bright contrast in stroboSCAT in Fig. 1E: The probe at 1.55 eV is initially pre-resonant with the exciton transition (dark contrast) but switches to resonant with the induced polaron absorption after the red- shift, generating a bright contrast (additional stroboSCAT datasets at different probe wave- lengths are displayed in fig. S13). Below, we discuss measurements of correlated spatio- energetic dynamics, fluence-dependent red- shift dynamics, and population saturation that cement our assignment of these redshift dy- namics to polaron formation. Polarons should exhibit correlated spatial and spectral dynamics, given that energetic stabi- lization associated with polaron formation results in a modification of transport proper- ties. To directly image these correlated dynamics and further support the correspondence be- tween the redshift timescale and the polaron formation timescale, we developed an approach capable of simultaneously resolving spectral and spatial evolution by merging stroboSCAT with transient reflectance microscopy (fig. S2). Figure 2C displays spatially resolved transient spectra at different pump-probe time delays (see also movies S2 and S3). At early times, the transient reflectance signal associated with ex- citons is concentrated around the pump exci- tation location at 0 mm. Between 4 and 50 ps, a V-shaped pattern emerges (highlighted with arrows in Fig. 2C), indicating that the spectral redshift associated with polaron formation is correlated to transport away from the excitation location. By 1 ns, only the redshifted polaron spec- tral signature remains, showing propagation over several micrometers. The correlated spatio- energetic dynamics are a clear indication that Tulyagankhodjaev et al., Science 382, 438–442 (2023) 27 October 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Spatio-energetic tracking of polaron formation in Re6Se8Cl2. (A) Transient reflectance spectra near the band edge after 2.41 eV pump excitation at an initial exciton density of 5 × 1017/cm3. DR/R refers to the pump- induced change in reflected light intensity. (B) Trace of the redshift highlighted with a dashed line in (A). The red line is a fit using a damped oscillator model with a frequency of 0.31 THz and damping time of 0.37 ps, combined with a 3-ps exponential decay. (C) Spacetime-transient reflection spectra after 2.41 eV pump excitation at an initial exciton density of 8 × 1018/cm3 for three time delays. The correlated redshift and transport dynamics are emphasized with arrows in the middle panel. Movies S2 and S3 display these coupled spatio-energetic dynamics for different excitation fluences. Data acquired at 295 K. excitons in Re6Se8Cl2 become substantially mo- bile only after they have formed exciton-polarons; bare excitons are effectively immobile, whereas polarons propagate over micrometers. The ob- served transport dynamics also rule out traditional exciton self-trapping (small polarons), which would reduce the exciton mobility upon polaron formation (32). These results illustrate the power of our correlative approach for understanding how polaron formation affects transport. Finally, we confirmed the formation of pola- rons in Re6Se8Cl2 and experimentally inferred their size by determining the density at which they begin to interact. At high densities, lattice deformations compete to displace the same atoms, resulting in a diminished ability to form polarons (Fig. 3A) (32). stroboSCAT data dis- played in Fig. 3B shows that as the photo- excitation density increases, the bare exciton signal (dark contrast) begins to dominate over the polaron signal (bright contrast), indicating that polaron formation is suppressed in re- gions of high excitation density. Figure 3C compares the polaron population (red trace) and bare exciton population (black trace) as a function of excitation fluence (analysis in fig. S14). We observed clear saturation of the polaron population at exciton densities between 0.45 × 1018/cm3 and 1.8 × 1018/cm3 (highlighted with a blue rectangle). This behavior signals that polarons are overlapping in space, analo- gous to a Mott transition that prevents further polaron formation (33). The onset of polaron saturation is also reflected in a transition from superdiffusive to almost diffusive transport (Fig. 3D), which we attribute to polaron–polaron scat- tering above the saturation density. The sup- pression of exciton-to-polaron conversion above saturation is most evident in the spectral dy- namics (Fig. 3E and fig. S15), where the redshift time associated with exciton stabilization in- creases from ~1.5 ps to hundreds of picoseconds. We reproduced these spectral dynamics (right panels of Fig. 3E) with a saturation model ac- counting for a kinetic blockade and polaron transport away from the excitation area (figs. S15 and S16). According to the Mott criterion (34), the polaron interaction radius associated with the observed critical density range of 0.45 × 1018/ cm3 to 1.8 × 1018/cm3 is 2.1 to 3.4 nm, corre- sponding to three- to five-unit cells in Re6Se8Cl2. Acoustic polarons are responsible for wavelike transport Large polarons, known to form in materials such as lead halide perovskite semiconductors, have been suggested to partially shield carrier- lattice scattering (35). Nevertheless, experiments consistently demonstrate a diffusive transport regime (10, 18, 19, 36), with sub-100-fs scat- tering times that indicate insufficient shielding to switch into the much-desired macroscopic ballistic transport regime. In contrast, the sustained quasi-ballistic behavior observed in Re6Se8Cl2 is reminiscent of acoustic polarons, which can form in low-dimensional materials and were theoretically invoked to rationalize the transport properties of polydiacetylene in one dimension (37, 38). The formation of acoustic polarons is rare, as is our observation of micrometer-scale exci- ton mean free paths at room temperature. To rationalize this notable behavior in Re6Se8Cl2, we use an approximate strong-coupling theory that describes an exciton of mass m coupled to acoustic phonons with a strength quantified by the deformation potential D in two dimensions. The acoustic phonons derive from superatoms of mass M with intercluster vibrational fre- quency W. In two dimensions, the energy to form a circular polaron with radius a and area pa2 is E ¼ ħ2 2ma2 þ 1 2 MW2D2pa2 (cid:2) DD ð1Þ where D is the dimensionless lattice displace- ment and ħ is the reduced Planck constant (39). In this simple picture, the polaron is only bound if the electron–phonon interaction outweighs the energetic penalties associated with exciton localization and lattice deformation. Minimizing Eq. 1 with respect to the lattice displacement D gives the existence criterion l > lc; l ¼ mD2 2ħ2MW2 = 2 ≈ 1:6 ¼ D2 4JMs2 ; ð2Þ lc ¼ p where l is a dimensionless measure of the exciton–acoustic phonon coupling strength, and in the second equality we have intro- duced the exciton transfer integral J and the speed of sound s = Wa. Applying density functional theory to Re6Se8Cl2, we calculate D = 4.4 eV (13). Taking into account the ex- perimentally inferred electronic band flat- tening due to intercluster optical phonons, which increases the exciton effective mass from 1.9me at 0 K to 60me at 300 K (7), we calculate a very small J of 1.5 meV at 300 K (13). When combined with the other material parameters (table S3), we find that l = 7 > lc, predicting a strongly bound polaron. Figure 4A plots the coupling strength l for Re6Se8Cl2, monolayer WSe2, crystalline pentacene, and 2D organic-inorganic halide perovskites. Re6Se8Cl2 has a coupling strength l that is 10 to 1000 times greater than that of the other materials and is thus the only material considered in Fig. 4 that is predicted to exhibit bound acous- tic polarons according to the criterion in Eq. 2. Within this theory, the key parameters setting Re6Se8Cl2 apart from other 2D and 3D semi- conductors is the combination of a quasi-flat electronic band structure at room temperature (small J) and strong exciton–acoustic phonon interactions (large D), yielding a large coupling strength l and associated strongly bound acoustic polarons. The polaron binding energy is reduced with decreasing temperature owing to the increase in the transfer integral J (7, 13). The estimate in Eq. 2 predicts that the polaron is not bound below ~175 K. Temperature-dependent stroboSCAT experiments (fig. S17) display a drastic reduction in MSD below ~150 K, lend- ing support to our central hypothesis and theory of acoustic polarons. We emphasize that Eq. 2 only provides a qualitative criterion for polaron formation and a detailed under- standing of the polaron stability and lifetime at finite temperature requires a more com- plete theoretical treatment (13). Tulyagankhodjaev et al., Science 382, 438–442 (2023) 27 October 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Determining polaron size in Re6Se8Cl2. (A) Depiction of polaron blockade at high densities. (B) Femtosecond stroboSCAT images at pump/probe energies of 2.41 eV/1.55 eV at initial exciton densities ranging from 1 × 1018/cm3 to 20 × 1018/cm3 in a 350-nm-thick Re6Se8Cl2 flake on glass. Bare excitons are associated with dark contrast, whereas polarons are associated with bright contrast. The rings observed at high exciton densities are diffraction rings. All scale bars are 3 mm. (C) Relative populations of polarons and bare excitons at a pump-probe time delay of 5 ps as a function of initial exciton density, indicating saturation of the polaron population. The blue shaded region for exciton densities between 0.45 × 1018/cm3 and 1.8 × 1018/cm3 indicates the range of critical polaron overlap density. (D) MSD exponent, a, as a function of initial exciton density. Error bars are 1 SD. (E) Experimental (left) and simulated (right) transient reflectance spectra taken at the center of the focused pump excitation (B) for different initial exciton densities. Simulations are based on a saturation model accounting for exciton-to-polaron conversion and polaron transport away from the excitation spot (13). Data acquired at 295 K. Fig. 4. Energy of an acoustic polaron. (A) The coupling strength l for different systems, presented as a function of their exciton transfer integral J. The boxed regions represent the upper and lower bounds for a range of deformation potentials and exciton transfer integrals within 30% of reported data (material properties collected in table S3). The critical value of lc = 1.6 is plotted as a dashed line. (B) Calculated polaron dispersion for Re6Se8Cl2 (13). The band effective mass is estimated by a parabolic fit at the minimum (dashed blue line). The solid red line emphasizes the linearity of the dispersion for higher values of momentum. Rationalizing the quasi-ballistic dynamics of acoustic polarons requires a more sophis- ticated quantum mechanical treatment. We generalize the stationary polaron description above and propose a variational wave function for the moving polaron defined by its average crystal momentum (13). Energy minimization produces the polaron dispersion shown in Fig. 4B. Near the band bottom, we extract a large effective polaron mass m* ≈ 200me, a substantial increase from the bare exciton mass of 60me at 300 K. More importantly, at higher momenta, the polaron inherits the linear dispersion of acoustic phonons—a renormalization evocative of light–matter hybridization to form polari- tons (40). The linear dispersion of acoustic polarons with a slope below that of other acous- tic phonons implies weak scattering because there are no dissipation channels that conserve both energy and momentum (38). The polaron is thus predicted to move quasi-ballistically at a speed proportional to the speed of sound of the lattice, consistent with our experimen- tal observations. Discussion and outlook We have observed a transport regime medi- ated by acoustic exciton-polarons in the vdW superatomic semiconductor Re6Se8Cl2. Polaron formation shields excitons from scattering with lattice phonons, resulting in quasi-ballistic elec- tronic energy flow over several micrometers within a nanosecond at room temperature. We reveal a very long exciton mean free path of ~1 mm, suggesting the possibility of ballistic excitonic transistors. Our discovery of this regime in a material with weak electronic dispersion provides an alternative to the cur- rent paradigm of increasing electronic conju- gation to improve transport. Indeed, our model for 2D acoustic polarons suggests that quasi-flat electronic bands and strong electron–phonon interactions can counterintuitively result in Tulyagankhodjaev et al., Science 382, 438–442 (2023) 27 October 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E exceptional electronic transport. Beyond 2D superatomic materials such as Re6Se8Cl2 and the recently reported graphullerene (41), moiré su- perlattices of 2D semiconductors may provide an interesting testing ground for acoustic polarons. Their superlattice potentials enable tuning elec- tronic bands (42, 43) to achieve values of J down to ~0. Combined with their strong deformation potentials (44), these flat bands could yield acoustic polarons with tunable transport prop- erties across a large temperature range. Gen- eralizing wavelike, ultralong-range electronic energy flow in 2D materials could herald an era of essentially lossless nanoelectronics. RE FE RENCES AND N OT ES 1. A. Javey, J. Guo, Q. Wang, M. Lundstrom, H. Dai, Nature 424, 654–657 (2003). 2. G. D. Scholes et al., Nature 543, 647–656 (2017). 3. X. Zhong et al., Nano Lett. 18, 1483–1488 (2018). 4. B. Choi et al., J. Am. Chem. Soc. 140, 9369–9373 (2018). 5. E. A. Doud et al., Nat. Rev. Mater. 5, 371–387 (2020). 6. E. J. Telford et al., Nano Lett. 20, 1718–1724 (2020). 7. Q. Li, F. Liu, J. C. Russell, X. Roy, X. Zhu, J. Chem. Phys. 152, 171101 (2020). 8. K. Lee et al., Adv. Mater. 31, e1903209 (2019). 9. 10. M. Delor, H. L. Weaver, Q. Yu, N. S. Ginsberg, Nat. Mater. 19, J. Chen, A. J. Millis, D. R. Reichman, Phys. Rev. Mater. 2, 114801 (2018). 56–62 (2020). 11. T. Zhu, J. M. Snaider, L. Yuan, L. Huang, Annu. Rev. Phys. Chem. 70, 219–244 (2019). 12. N. S. Ginsberg, W. A. Tisdale, Annu. Rev. Phys. Chem. 71, 1–30 (2020). 13. Materials and methods are available as supplementary materials. 14. C. Jacoboni, C. Canali, G. Ottaviani, A. Alberigi Quaranta, Solid-State Electron. 20, 77–89 (1977). 15. J. Shin et al., Science 377, 437–440 (2022). 16. S. Yue et al., Science 377, 433–436 (2022). 17. G. M. Akselrod et al., Nat. Commun. 5, 3646 (2014). 18. J. Sung et al., Nat. Phys. 16, 171–176 (2020). 19. Z. Guo et al., Science 356, 59–62 (2017). 20. M. M. Glazov, Phys. Rev. B 100, 045426 (2019). 21. L. M. Smith et al., Phys. Rev. B 39, 1862–1870 (1989). 22. M. Kulig et al., Phys. Rev. Lett. 120, 207401 (2018). 23. R. Perea-Causín et al., Nano Lett. 19, 7317–7323 (2019). 24. P. Ruello, V. E. Gusev, Ultrasonics 56, 21–35 (2015). 25. J. Rudolph, R. Hey, P. V. Santos, Phys. Rev. Lett. 99, 047602 (2007). 26. E. Baldini et al., Sci. Adv. 5, eaax2937 (2019). 27. K. Datta et al., Nat. Photonics 16, 242–247 (2022). 28. K. Miyata et al., Sci. Adv. 3, e1701217 (2017). 29. F. X. Morrissey, J. G. Mance, A. D. Van Pelt, S. L. Dexheimer, J. Phys. Condens. Matter 25, 144204 (2013). 30. B. Guzelturk et al., Nat. Mater. 20, 618–623 (2021). 31. B. Wu et al., Sci. Adv. 7, eabd3160 (2021). 32. D. Emin, Polarons (Cambridge Univ. Press, 2012). 33. J. M. Frost, L. D. Whalley, A. Walsh, ACS Energy Lett. 2, 2647–2652 (2017). 34. P. P. Edwards, M. J. Sienko, Phys. Rev. B 17, 2575–2581 (1978). 35. H. Zhu et al., Science 353, 1409–1413 (2016). 36. Z. Guo, J. S. Manser, Y. Wan, P. V. Kamat, L. Huang, Nat. Commun. 6, 7471–7479 (2015). 37. H.-B. Schüttler, T. Holstein, Phys. Rev. Lett. 51, 2337–2340 (1983). 38. E. G. Wilson, J. Phys. C Solid State Phys. 16, 6739–6755 (1983). 39. G. Whitfield, P. B. Shaw, Phys. Rev. B 14, 3346–3355 (1976). 40. D. L. Mills, E. Burstein, Rep. Prog. Phys. 37, 817–926 (1974). 41. E. Meirzadeh et al., Nature 613, 71–76 (2023). 42. L. Wang et al., Nat. Mater. 19, 861–866 (2020). 43. F. Wu, T. Lovorn, E. Tutuc, A. H. MacDonald, Phys. Rev. Lett. 121, 026402 (2018). 44. D. Waters et al., ACS Nano 14, 7564–7573 (2020). 45. J. Tulyagankhodjaev et al., Dataset for “Room temperature wavelike exciton transport in a van der Waals superatomic semiconductor,” version 1, Zenodo (2023); https://doi.org/10.5281/zenodo.8183811. ACKN OWLED GMEN TS We are grateful to P. Batail for bringing this family of materials to our attention and for ensuing productive discussions. We also thank L. E. Brus, X. Zhu, C. Nuckolls, and D. R. Reichman for helpful discussions and D. Xu and J. Baxter for technical help with measurements and analysis. Funding: This material is primarily based on work supported by the National Science Foundation (NSF) through the Columbia MRSEC on Precision-Assembled Quantum Materials (PAQM) (DMR-2011738) (M.D., X.R., and T.C.B.) and by the Air Force Office of Scientific Research (AFOSR) under grant number FA9550-22-1-0389 (M.D. and X.R.). stroboSCAT instrument development was supported by the NSF under grant number DMR-2115625 (M.D.). M.D. acknowledges support from the Arnold and Mabel Beckman Foundation through a Beckman Young Investigator award. X.R. and J.Y. acknowledge support from NSF CAREER Award DMR-1751949. J.A.T. was supported by an NSF Graduate Research Fellowship. J.C.R. was supported by a Department of Defense National Defense Science and Engineering Graduate Fellowship. Author contributions: J.A.T. and M.D. conceived of and designed the experiments. J.A.T. performed, analyzed, and simulated stroboSCAT and spacetime transient reflection experiments with assistance from A.C.S. J.Y., J.C.R., D.G.C., and X.R. synthesized and characterized Re6Se8Cl2 single crystals. J.Y. performed and analyzed heat capacity measurements. J.A.T., M.E.R., J.Y., and H.S. prepared exfoliated samples. P.S. and T.C.B. developed the theory and performed density functional theory calculations and polaron dispersion calculations. M.D. supervised the project. J.A.T., P.S., J.Y., X.R., T.C.B., and M.D. wrote the manuscript, with input from all authors. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data and code necessary to reproduce the figures in the main text and supplementary materials are available on Zenodo (45). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf2698 Materials and Methods Supplementary Text Figs. S1 to S17 Tables S1 to S3 References (46–71) Movies S1 to S3 Submitted 11 October 2022; accepted 21 September 2023 10.1126/science.adf2698 Tulyagankhodjaev et al., Science 382, 438–442 (2023) 27 October 2023 5 of 5
10.1126_science.adf2040
RES EARCH 2D MATERIALS A quantum ruler for orbital magnetism in moiré quantum matter M. R. Slot1,2†, Y. Maximenko1,3†, P. M. Haney1, S. Kim1,4, D. T. Walkup1, E. Strelcov1,3, Son T. Le1, E. M. Shih1,3, D. Yildiz1,4, S. R. Blankenship1, K. Watanabe5, T. Taniguchi6, Y. Barlas7, N. B. Zhitenev1, F. Ghahari8*, J. A. Stroscio1* For almost a century, magnetic oscillations have been a powerful “quantum ruler” for measuring Fermi surface topology. In this study, we used Landau-level spectroscopy to unravel the energy-resolved valley-contrasting orbital magnetism and large orbital magnetic susceptibility that contribute to the energies of Landau levels of twisted double-bilayer graphene. These orbital magnetism effects led to substantial deviations from the standard Onsager relation, which manifested as a breakdown in scaling of Landau-level orbits. These substantial magnetic responses emerged from the nontrivial quantum geometry of the electronic structure and the large length scale of the moiré lattice potential. Going beyond traditional measurements, Landau-level spectroscopy performed with a scanning tunneling microscope offers a complete quantum ruler that resolves the full energy dependence of orbital magnetic properties in moiré quantum matter. M oiré quantum matter (MQM) systems (1) consist of stacked and twisted layers of van der Waals materials and have emerged as versatile condensed-matter quantum simulators (2). The twist angle, choice of material, number of layers, and the application of electric and magnetic fields provide a vast arena for realizing quantum phases that result from the interplay between electron correlations and topology. The dis- covery of flat electronic bands that host super- conductivity and correlated insulating states in magic-angle twisted bilayer graphene (MATBG) inspired rapid exploration of the parameter space (3–5). The most recent endeavors have focused on heterostructures of three, four, and five layers of alternating twisted graphene monolayers (6–11). The moiré systems of MATBG and related heterostructures support topological bands with nonzero Chern number (4, 12), which is de- rived from the Berry curvature of the Bloch wave functions (13). Berry curvature is intimately related to orbital magnetization, and indeed, previous work on MATBG and related systems has observed orbital magnetic order, valley Hall, 1Physical Measurement Laboratory, National Institute of Standards and Technology, Gaithersburg, MD 20899, USA. 2Department of Physics, Georgetown University, Washington, DC 20007, USA. 3Department of Chemistry and Biochemistry, University of Maryland, College Park, MD 20742, USA. 4Joint Quantum Institute, Department of Physics, University of Maryland, College Park, MD 20742, USA. 5Research Center for Functional Materials, National Institute for Materials Science, 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan. 6International Center for Materials Nanoarchitectonics, National Institute for Materials Science, 1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan. 7Department of Physics, University of Nevada, Reno, NV 89557, USA. 8Department of Physics and Astronomy, George Mason University, Fairfax, VA 22030, USA. *Corresponding author. Email: fghahari@gmu.edu (F.G.); joseph.stroscio@nist.gov (J.A.S.) †These authors contributed equally to this work. and quantum anomalous Hall effects (14–16). Berry curvature and the related quantum met- ric also provide significant contributions to the orbital magnetic susceptibility (17, 18). The valley-contrasting orbital magnetism and or- bital magnetic susceptibility play a central role in the material response to applied magnetic fields. Landau-level (LL) spectroscopy is a well- established tool to experimentally deduce the zero-field properties of the band structure, and in this work, we expand its application to extracting higher-order magnetic response functions. Semiclassically, LLs are described by the Onsager relation, relating the extremal cross- sectional area of the Fermi surface to the pe- riod of oscillations in the de Haas–van Alphen effect (19). The orbital magnetism and orbital magnetic susceptibility give rise to energy- dependent first- and second-order corrections with magnetic field in the Onsager relation (20), which was recently formulated as a general expansion in higher-order response functions (21). To determine these effects and their full energy dependence, high-resolution measurements of the LLs in the electronic bands are required: a “quantum ruler” prob- ing the evolution of LLs at all energies as a function of displacement field and magnetic field. In this work, we used LL measurements as a quantum ruler for twisted double-bilayer graph- ene (TDBG) to quantitatively determine the tunable electronic structure, energy-dependent valley-contrasting orbital magnetism, and or- bital magnetic susceptibility. TDBG is MQM with highly reconfigurable bands; the Bernal bilayer constituents have an electrostatically tunable band structure in themselves (22–31). Flat bands exist over a wide range of small twist angles (q), from about 0.8° to 1.5°. Pre- vious transport and scanning tunneling micros- copy (STM) measurements have focused on correlated states, which form at various partial fillings (22–30), or density wave states at larger twist angles of q ≈ 2.4° (31). Our intermediate twist angle of 1.74° gives rise to moiré mini- bands that are narrow but not extremely flat. This slightly wider bandwidth allows LLs to be resolved over the entire band structure, which has not been observed in previous local probe measurements (29). Tuning the electronic structure of TDBG Figure 1A shows the simulated band structure of TDBG at q = 1.75° with the Bistritzer– MacDonald (BM) continuum model (3). The moiré periodic potential leads to narrow moiré minibands, composed of two low-energy bands—the valence V1 and conduction C1 bands—and higher remote energy bands—V2 and C2. The narrow bands V1 and C1 have a bandwidth of about 50 meV and are isolated from other bands (Fig. 1B). The TDBG system displays a large tunability in its electronic structure (12, 32–39). A perpendicular electric field induces a potential difference between the layers of TDBG, drastically altering the V1 band structure. The V1 band is relatively flat at zero displacement field (D) (Fig. 1B) and de- velops pronounced electron and hole pockets at D ≠ 0 (Fig. 1C and fig. S8). In the next section, we first describe the measurement of this band structure evolution under a dis- placement field. Measurements were performed in a custom- built dilution refrigerator–based scanning probe system operating at 10 mK (40, 41). Figure 1D shows the STM topograph of the TDBG sam- ple used in this study (fig. S1). The measured moiré periodicity of l = (8.1 ± 0.1) nm cor- responds to a twist angle of q = (1.74° ± 0.02°), and the local heterostrain was found to be <0.1%, as determined from atomically resolved spatial measurements (42). Scanning tunnel- ing spectroscopy (STS) measures the differen- tial tunneling conductance (dI/dV) signal, which is proportional to the local density of states (LDOS) (43). For TDBG, the measured dI/dV reflects the LDOS of the topmost layer of the four-layer system. The finite tip potential leads to a fixed top gate VT, determined by the contact-potential difference between probe and sample. Varying the back-gate voltage VG simultaneously tunes the displacement field D and the carrier density n following D º CGVG − CTVT and n º CGVG + CTVT, where CG and VG are the back-gate capacitance and voltage, and CT and VT are tip capacitance and potential, respectively (42). We mapped the differential tunneling conductance as a function of sample bias VB, the displacement field D, carrier density n, and magnetic field B. Figure 1E shows the experimental LDOS map at B = 0 T. The low-energy bands V1 and C1 and remote bands V2 and C2 and their Slot et al., Science 382, 81–87 (2023) 6 October 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E A D 18.8 U = 0 meV U = 0 meV U = 80 meV C2 C1 V1 V2 B ) V e m ( E 150 100 50 0 -50 -100 -150 C ) V e m ( E 150 100 50 0 -50 -100 -150 C2 C1 EF V1 V2 hole pocket C2 C1 EF electron pocket V1 V2 saddle point g k' m k g g k' m k g Z (pm) 108.8 E Experimental dI/dV (nS) 2 4 B = 0 T 0 F B = 0 T 0 1 Theory LDOS (arb. units) 2 3 4 5 6 0.4 - ) 1 m n V ( D 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 V1 C1 C2 0.2 saddle point V2 V1 C1 0 n / n 0 -0.2 -0.4 -0.6 - ) 1 m n V ( D 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 0.4 V1 C1 C2 0.2 saddle point V2 V1 C1 0 n / n 0 -0.2 -0.4 -0.6 1 nm 10 nm -100 -50 0 (mV) V B 50 100 -100 -50 0 E (meV) 50 100 Fig. 1. Tuning the narrow electronic bands in TDBG with displacement field at q = 1.75°. (A) Energy-band structure of TDBG showing the four low-energy bands computed by using a continuum model with interlayer potential U = 0 meV and q = 1.75°. Line cuts from the energy bands in (A) along Bloch wave vector kx = 0 for (B) U = 0 meV and (C) U = 80 meV. A dashed horizontal line indicates the position of the Fermi level (EF), which moves through the bands as the displacement field and carrier density are varied. The x-axis labels (g, k′, m, and k) denote high-symmetry points in the mini-Brillouin zone. (D) STM topographic image of the TDBG moiré pattern, with moiré wavelength l = (8.1 ± 0.1) nm. (Inset) A magnified portion of the image showing the atomic lattice of graphene. A twist angle of q = (1.74° ± 0.02°) was determined from the atomically resolved moiré lattice (42). The red dot indicates the ABBC symmetry position where the spectral maps with applied magnetic field were acquired. Tunneling setpoint I = 20 pA; VB = 100 mV; temperature T = 0.01 K. (E) Tunneling spectral maps in the D versus VB plane for B = 0 T. n0 (–n0) indicates the electron density at full filling of the C1 (V1) band. Setpoints: I = 20 pA, VB = 120 mV, T = 0.01 K. (F) Corresponding theory map for B = 0 T (continuum-model calculation). arb. units, arbitrary units. evolution with displacement field are clearly visible. The experimental data qualitatively agree with the modeled LDOS of the top layer of the four-layer graphene system obtained with continuum-model calculations [Fig. 1F (42)]. The calculations in this work are pre- sented for qualitative comparison and are not meant for quantitative comparison with the experiment. The trends in the dI/dV map can be quali- tatively understood by considering how the theoretically predicted bands in Fig. 1, B and C, fill with n as the Fermi level is raised from negative to positive energies while the bands are modified by the displacement field D. At negative displacement field approaching D = 0, the Fermi level is initially pinned to the flat portions of V1 with its associated high LDOS (Fig. 1B). This appears in the spectral maps (Fig. 1, E and F) with the high LDOS of the V1 band located near the Fermi level, correspond- ing to zero sample bias. Above a displacement field of 0.2 V/nm, the relatively flat C1 band becomes pinned at the Fermi level, and the V1 band develops a larger dispersion while its highest LDOS at its saddle point moves to lower energies away from C1 with increasing positive displacement field (Fig. 1C). This re- sults in a bright spectral line associated with the saddle point in the V1 band with a negative slope in the top portion of the spectral maps in Fig. 1, E and F. Mapping Landau levels Application of a magnetic field creates sharp LLs in the dI/dV spectrum of all four bands, as shown in the dI/dV maps at B = 4 T and 8 T in Fig. 2, B and D, respectively (see figs. S11 to S13 for additional data sets). All dI/dV maps with magnetic field are acquired on an ABBC site (Fig. 1D, red dot). The experimental maps can be compared with the theoretical maps in Fig. 2, C and E, obtained from magnetic field–dependent quantum mechanical calculations (42), follow- ing the methods in (44). A series of LL spectra at different displacement fields are extracted from the experimental map at B = 4 T (Fig. 2A). The LLs in the V1 band are clearly pronounced and show that the LL energies are irregularly spaced, showing neither the equal spacing char- acteristic for a parabolic band nor the square root dependence on orbital index as observed in graphene. In the next section, we use the evolution of the V1 LL spacing to experimentally demonstrate the characteristics of the tunable band. The electron-like versus hole-like character of a band is manifested in the energy variation with LL index and magnetic field B. LLs orig- inating from electron-like Fermi pockets in- crease in energy with LL index and B, whereas the opposite trend applies for hole-like LLs. The saddle-point energy separates electron-like from hole-like pockets. With this in mind, we examined the experimental spectra of Fig. 2A. At negative displacement field, a set of well- resolved V1 LLs is clearly visible in the dI/dV map at B = 4 T (Fig. 2B) and the corresponding extracted spectra (Fig. 2A). We identified the zeroth LL at the bottom of the V1 band, which was at VB ≈ –45 mV for zero displacement field. For a given displacement field in Fig. 2A, the LL energy spacing in the V1 band is ob- served to gradually decrease with increasing bias voltage, followed by a gradual increase, Slot et al., Science 382, 81–87 (2023) 6 October 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E A V2 B = 4 T V1 hole pocket C1 B B = 4 T 0 dI/dV (nS) 2 1 3 ) s t i n u . b r a ( V d / I d 2.5 2 1.5 1 0.5 0 electron pocket D = -0.05 V nm-1 D = -0.10 V nm-1 hole pocket electron pocket saddle point 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 ) 1 - m n V ( D C ) n / n 0 1 - m n V ( D B = 4 T 0 -0.2 LDOS (arb. units) 0.2 0.6 0.4 0.8 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 0 (mV) V B dI/dV (nS) 2 1 D = -0.15 V nm-1 -100 -50 D = -0.20 V nm-1 D = -0.25 V nm-1 electron pocket hole pocket B = 8 T 0 D 0.5 ) 1 - m n V ( D 0.4 0.3 0.2 0.1 0 -0.1 -0.2 50 100 -100 -50 0 E (meV) 50 E ) n / n 0 1 - m n V ( D 3 0.4 0.2 0 -0.2 -0.4 -0.6 -0.8 B = 8 T 0 -0.2 LDOS (arb. units) 0.2 0.6 0.4 0.8 0.5 0.4 0.3 0.2 0.1 0 -0.1 -0.2 0.4 0.2 0 -0.2 n / n 0 -0.4 -0.6 -0.8 100 0.4 0.2 0 -0.2 n / n 0 -0.4 -0.6 -0.8 100 -100 -50 0 50 -100 -50 V B (mV) 0 (mV) V B 50 100 -100 -50 0 E (meV) 50 Fig. 2. Tuning LLs in TDBG with displacement and magnetic fields. (A) Tunneling spectra versus VB of TDBG at selected displacement fields taken from the map in (B). Tunneling spectral map in the D versus VB plane at (B) B = 4 T. (C) Corresponding theory map for B = 4 T (quantum calculation). (D) Tunneling spectral map in the D versus VB plane at B = 8 T. (E) Corresponding theory map for B = 8 T (quantum calculation) (42). The maps in (B) and (D) were acquired on the maxima positions in the moiré structure indicated by the red dot in Fig. 1D. Setpoints: I = 20 pA, VB = 120 mV for (B) and (D), T = 0.01 K. See figs. S11 to S13 for additional spectral map data. pointing to a transition from an electron-like to a hole-like pocket through a saddle point. The LL spacing is inversely proportional to the LDOS, which peaks at the Van Hove sin- gularity at the saddle point. The saddle point is pinned at the Fermi level for negative carrier densities and visible as the high LDOS cross- ing through the V1 band at positive densities (Fig. 2, B and D). We used the B-field dependence of the LLs to confirm the nature of the tunable band struc- ture. Figure 3, A to C, displays dI/dV maps in the B versus VB planes at a fixed displacement field, showing the dispersion of the LLs with B in the V1 band. At negative displacement fields (Fig. 3, B and C), the LLs below the saddle point at the Fermi level disperse with positive slope, and those above the saddle point disperse with negative slope. This con- firms the electron versus hole-pocket nature of the V1 band. The transition from electron to hole pocket in the V1 band is similar at positive displacement fields, where increasing displacement fields tunes the V1 saddle point to lower energies (Fig. 3A). At D = 0.35 V nm–1 (Fig. 3A), the saddle point is tuned to the mid- dle of the band with an equal number of posi- tive and negative dispersing LLs on either side, which demonstrates the well-determined tun- ability of electron and hole pockets with dis- placement field. Deviations from the Onsager relation We now zoom in on the detailed features in the V1 band, leveraging LL spectroscopy to extract quantitative information about the zero-field material properties involving valley- contrasting orbital magnetism and orbital mag- netic susceptibility. Figure 3D shows extracted, well-resolved spectra at selected B fields. In Fig. 3E, we show the corresponding peak po- sitions of the LLs in the V1 band versus B for n = 0 to 7, which shows how the LLs disperse with magnetic field. LLs are semiclassically described by Onsager’s quantization condition, which requires that the total phase accumula- tion over a cyclotron orbit be an integer mul- tiple of 2p (19). This yields S Enð Þf 0 =2p ¼ 2pBn n þ 1=2 ð Þ ð1Þ relating the k-space area S(En) of the zero-field iso-energy contour of the nth LL En to the magnetic field Bn, where f0 = h/e is the mag- netic flux quantum, h is Planck’s constant, ½ and e is the elementary charge. With the equal- energy k-space area scaling as Bn(n + 1/2) for all LLs at energy En, a plot of Sº Bn n þð 1=2Þ(cid:2) n¼0;1;2… versus LL energy for all LLs would collapse onto a single curve. From Fig. 3E, we plot Sº½Bnðn þ 1=2Þ(cid:2) n¼0…7 in Fig. 3F. The fact that experimental data points fail to collapse onto a single curve demonstrates the in- adequacy of the Onsager relation and the need for higher-order corrections. Recently, a systematic expansion of the semi- classical Onsager relation described in Eq. 1 was derived by adding corrections in terms of powers of B (21): Bn n þ 1=2 ð m′ Enð Þ=f0 ¼ S Enð ÞB2 n Þ=4p2 þ =2 þ … ÞBn þ c′ Enð ð2Þ The coefficients of the first- and second-order correction, m′(E) and c′(E), are the derivatives with respect to energy of the total orbital mag- netic moment and orbital magnetic susceptibil- ity, respectively. Eq. 2 is thus a powerful tool to either infer the zero-field material properties S(E), m′(E), and c′(E) from a given LL spec- trum, or to predict the detailed LL spectrum given the knowledge of the zero-field material properties. In the next section, we extract these properties for TDBG from the LL spectrum Slot et al., Science 382, 81–87 (2023) 6 October 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E electron pocket saddle point hole pocket D = 0.35 V nm-1 A ) T ( B 8 6 4 D -60 -40 -20 V B 20 0 (mV) 40 60 saddle point electron pocket hole pocket D = -0.05 V nm-1 B ) T ( B 8 6 4 -60 -40 -20 V B 20 0 (mV) 40 60 saddle point hole pocket D = -0.10 V nm-1 electron pocket C ) T ( B 8 6 4 -60 -40 -20 V B 20 0 (mV) 40 60 D ) s t i n u . b r a ( V d / I d 2 1 0 -100 LL0 LL0 V1 band LL0 D = -0.05 V nm-1 B = 8 T B = 6 T B = 4 T -50 0 B (mV) V 50 100 D = -0.05 V nm-1 E ) T ( B 12 10 8 6 4 LL0 2 -50 LL1 LL2 LL3 -40 -30 V -20 B (mV) -10 0 n=0 n=1 n=2 n=3 n=4 n=5 n=6 n=7 D = -0.05 V nm-1 0.4 F 0.35 ) 2 - m n ( S 0.3 0.25 0.2 0.15 0.1 0.05 low dI/dV (arb. units) high 0 -50 -40 S(E) to 0th Order -30 -20 E (meV) -10 0 Fig. 3. Breakdown of the standard Onsager relation in TDBG. Spectral maps in the B versus VB plane for displacement fields of (A) D = 0.35 V nm–1, (B) D = –0.05 V nm–1, and (C) D = –0.1 V nm–1 showing the dispersion of the electron and hole-pocket LLs with magnetic field, and the shift of the V1 saddle point to lower energies with increasing displacement field. (D) Tunneling spectra versus VB of TDBG as a function of magnetic field at fixed D = –0.05 V nm–1. (E) LL peak positions versus B for D = –0.05 V nm–1. Data from LLs n = 0 to n = 8 are shown. Symbols, experimental peak positions; solid lines, cubic-spline interpolation of experimental peak positions. (F) S(E) computed from the solid lines in (E) with Eq. 1. The lack of data collapse indicates that second-order contributions to the LL energies are substantial. and show a large contribution of c′(E ) for moiré systems that scale with an increase in moiré lattice constant. Extracting zero-field properties The first-order correction m′(E) includes con- tributions from the total Berry curvature en- closed by an equal energy contour (Fig. 4, A and B), which yields the well-known Berry phase correction, and the average orbital mag- netic moment per carrier at energy E (21, 45). The calculated Berry curvature W for the V1 band is represented by a hue in the mini-Brillouin zone for each valley K and K′ in the Brillouin zone in Fig. 4, A and B, respectively. The valleys have equal and opposite Berry curva- ture and orbital moment. By symmetry, the total intravalley orbital magnetic moment is finite only at nonzero displacement field. This tunability of the valley-contrasting orbital mag- netism enables the disentangling of the orbital susceptibility from the orbital magnetic moment, which we use in the analysis of the LLs below. The strength of these contributions in Eq. 2 can be evaluated by using the continuum model. Figure 4C shows the terms on the right side of Eq. 2 for LL0, calculated with the zero-field continuum model and multiplied by f0/B to obtain the phase contribution in the extended Onsager relation in Eq. 2 (see fig. S5 for all phase contributions). We note that these phase terms are energy dependent and are large in TDBG; for comparison, a value of 0.5 would be equivalent to the p Berry phase in single-layer graphene. The sign of the orbital moment and its derivative m′(E) is opposite for the K and K′ valleys, as expected from the valley-contrasting Berry curvature shown in Fig. 4, A and B. Figure 4D shows how the phase contribu- tions displayed in Fig. 4C drastically alter the semiclassical LL spectrum. The semiclassical, zeroth-order LL spectrum (Eq. 1) is displayed in blue for the lowest LLs in the V1 band by using S(E) obtained from the zero-field contin- uum model. The first-order correction m′(E ) leads to a valley splitting of the LLs indicated by the dashed red and black lines. The calcu- lated valley splitting at the bottom of the V1 band is substantial but within the LL spacing, where the LL index can be assigned unambig- uously. This makes LL0 ideally suitable to de- termine the experimental valley splitting from our high-resolution measurements. High-resolution dI/dV spectra of LL0 at increasing magnetic fields 9 to 13 T at D = –0.25 V/nm are shown in Fig. 4E. We observed a clear experimental splitting, with the red lines indicating the fitted peak positions. At 13 T, the LL0 peak is very broad owing to many levels forming from the Hofstadter spectrum at high magnetic fields. We present a comparison to the spectrum computed quantum mechanically in Fig. 4F (gray peaks) to show the onset of Hofstadter minibands at higher fields. The values of splitting obtained with semiclassical calculations are similar (Fig. 4D). Using an energy broadening comparable to the experi- ment, shown as the black curve, we obtained qualitative agreement with the experiment. Figure 4G shows a direct comparison of the experimental LL0 splitting in Fig. 4E (blue data points) with the quantum mechanical calculations in Fig. 4F (open orange circles) as a function of B field. We found semiquanti- tative agreement between the calculated and measured valley-splitting values, demonstrat- ing the tunable valley-contrasting orbital mag- netism present in TDBG. Most notably, the splitting was much larger than the standard Zeeman effect with Landé g factor of 2 (solid black line) and did not increase linearly with B field as one would expect for a Zeeman- type splitting mechanism. The latter indicates that the splitting is not only given by m′(B), but rather, a considerable nonzero second- order correction c′ leads to nonlinearity in the splitting with B. We obtained an estimate of m′ in the middle of the V1 band of ≈ 3.5 × 1013 Slot et al., Science 382, 81–87 (2023) 6 October 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E A K Valley U = -40 meV k’ k m orb B K’ Valley U = -40 meV k’ m orb k Quantum K Quantum K' 0th order 1st order K 1st order K' 1st+2nd order K 1st+2nd order K' D 12 10 ) T ( B 8 6 4 2 E ) s t i n u . b r a ( V d / I d 13 T 12 T 11 T 10 T 9 T -20 -10 (mV) V B 0 F ) s t i n u . b r a ( S O D L 13 T 12 T 11 T 10 T 9 T G ) V e m ( g n i t t i l p s 0 L L 5.5 5 4.5 4 3.5 3 2.5 2 1.5 1 0.5 C n o i t u b i r t n o C e s a h P 0.2 0.15 0.1 0.05 0 -0.05 -0.1 -0.15 -0.2 -0.25 -0.3 -45 -40 -35 -30 -25 -20 E (meV) quantum U = -40 meV experiment D = -0.25 V nm-1 Zeeman g = 2.0 Zeeman g = 6.5 -40 -35 E (meV) -30 -40 -30 E (meV) -20 0 0 5 10 15 B (T) Fig. 4. LL splitting originating from valley-contrasting orbital magnetism and Berry curvature in TDBG. Calculated three-dimensional plot of the V1 band with Berry curvature shading at U = –40 meV for (A) K and (B) K′ valleys. K and K′ valleys have equal and opposite nonzero Berry curvature and equal and opposite orbital magnetic moment (orbital moments depicted schematically). Eb is the D-dependent energy of the bottom of the V1 band. (C) A comparison of the calculated first- (m′) and second-order (c′) correction magnitude with the extended Onsager relation for the V1 TDBG band (U = –40 meV). See fig. S5 and S6 for additional phase contribution results (42). (D) Calculation of the first few LLs (n = 0, 1, and 2) versus B, comparing quantum mechanical to semiclassical methods with zeroth-, first-, and second-order corrections. The first-order correction is grossly deficient in matching the quantum results, particularly in the n = 0 level, whereas including both first- and second-order corrections provides a good agreement with the quantum results (U = –40 meV). (E) Tunneling spectra of the LL0 peak as a function of B for D = –0.25 V nm–1 that show a splitting of the LLs in high magnetic fields. The red lines indicate peak positions determined from nonlinear least-square fitting of a double Lorentzian function to LL0. (F) Quantum calculation of LL0 as a function of magnetic field for an interlayer potential U = –40 meV, corresponding to a displacement field of D = –0.25 V nm–1. The LL shows a valley splitting upon application of a displacement field. The solid black lines have a Gaussian energy broadening of s = 1 meV to match experiment; gray lines, s = 0.05 meV. (G) The energy difference of the split n = 0 LL from (E) compared with quantum calculations in (F). For the quantum calculation in higher fields, we plot the splitting as the difference in energy between lowest levels in the Hofstadter bands of the K and K′ valleys. The uncertainty in the experimental points is derived from twice the standard deviation obtained from the nonlinear least-square fits of the peaks in (E). For comparison, the solid lines show the calculated Zeeman splitting with Landé g factor of (black) g = 2 and (red) 6.5. C/(J s), comparable to a calculated estimate of ≈ 4.4 × 1013 C/(J s), using the values for the splitting, S′(E) and c′(E), extracted from the ex- periment (42). The first-order correction m′ does not ex- plain the lack of collapse of the LLs onto a single curve. This is the result of a strong, experimentally observable second-order cor- rection, given by the energy-dependent orbital magnetic susceptibility c′(E). This contribu- tion originates from several mechanisms, as elu- cidated in recent theoretical works (18, 46, 47), including the geometrical origin, namely the Berry curvature and quantum metric (48). The calculated c′(E) of TDBG has the same sign for both the K and K′ valley, as shown for LL0 in Fig. 4C. This results in a shift of the valley-split LLs, quadratically increasing with B field, as shown by the black and red solid lines in the continuum-model calculation in Fig. 4D. This fully expanded semiclassical approxi- mation with first- and second-order correction (Eq. 2) agrees well with the quantum me- chanical spectrum obtained for the same system (red and black open circles), again emphasizing the importance of the second- order correction to adequately describe the system. For a large region of energies and mag- netic fields, the second-order correction is larger than the first one (Fig. 4, C and D, and fig. S6). This means that the energy-dependence of the magnetic field–induced orbital moment exceeds the displacement field–induced orbi- tal moment. To extract the second-order correction to the LL energy, we note that the first-order split- ting is typically not visually evident for most electron-like LLs in the V1 band at smaller displacement fields owing to smaller valley splitting from orbital magnetism, and pre- sumably, to broadening. The energy value of a single LL peak is then approximately the average of the valley-split values. The valley- averaged energy at a fixed displacement field is obtained from Eq. 2 by letting m′(E ) = 0, keeping only the second-order correction. For equal energy LLs n and m with associated fields Bn and Bm, Eq. 2 yields the following for S and c′: S Eð Þ ¼ 2pe h BnBm ð " Bm n þ 1=2 B2 m (cid:1) Þ (cid:3) Bn m þ 1=2 ð Þ (cid:3) (cid:3) B2 n # c′ Eð Þ ¼ e ph ð " Þ (cid:3) Bm m þ 1=2 Bn n þ 1=2 ð (cid:3) (cid:1) B2 n (cid:3) B2 m ð3Þ # Þ ð4Þ Figure 5A shows the average values of S(E) obtained by evaluating Eq. 3 for all pairs of LLs in Fig. 3E at each energy value. By contrast to Fig. 3F, S(E ) collapses onto a single curve (blue data points) in Fig. 5A, confirming the Slot et al., Science 382, 81–87 (2023) 6 October 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E experiment D = -0.05 V nm-2 theory U = -8 meV S(E) to 2nd Order A 0.4 ) 2 - m n ( S 0.35 0.3 0.25 0.2 0.15 0.1 0.05 B 10 5 0 -5 -10 -15 experiment D = -0.05 V nm-2 theory U = -8 meV 0 -50 -40 -30 -20 E (meV) -10 0 10 -20 -40 -30 -20 -10 0 10 E (meV) Fig. 5. Extraction of enhanced energy-dependent orbital magnetic susceptibility in moiré quantum matter from the extended Onsager relation. (A) The symbols represent S(E) computed from the solid lines in Fig. 3E with Eq. 3, demonstrating data collapse onto one continuous curve. The solid line represents S(E) computed for the V1 band from continuum-model calculations. (B) The symbols represent c′(E) computed from the solid lines in Fig. 3E with Eq. 4. The solid line represents c′(E) computed from continuum-model calculations. See (42) for a discussion of the error analysis. applicability of the second-order correction. The result has a similar trend to that of the zero-field continuum-model calculation of S(E) (orange solid line). Similarly, we obtained the average value of c′(E) by evaluating Eq. 4 (blue data points in Fig. 5B). The experimentally extracted values are in qualitative agreement with the zero- field orbital magnetic susceptibility (orange line), computed with the continuum model according to the formalism of (49). The un- certainty in the extracted c′(E ) and the de- viation in overlap of segments from different pairs of LLs are attributed to two factors. First, corrections to Landau quantization energies have a smaller impact on the high-index LLs located at higher energies (50). Extracting the relatively smaller corrections at these energies therefore leads to larger uncertainty. Second, variations of the chemical potential with mag- netic field shift the LL energy positions, espe- cially at high magnetic fields (42). We do not believe interactions make a qualitative impact on our analysis. The LL scaling obtained with the extended Onsager relation is highly con- strained, and we find it is well satisfied over a wide range of electron filling. Presumably, interactions would renormalize LL energies in a distinct manner, which would depend strongly on filling factor. Discussion and outlook There are several distinct microscopic mecha- nisms that underlie the orbital magnetic suscep- tibility, including recently identified geometric contributions (18, 46). In fig. S9, we show the contributions to c′(E) from these mechanisms [see (42) for the full breakdown]. The contri- butions from the Pauli paramagnetism and the quantum geometry involve the orbital mag- netic moment, Berry curvature, and quantum metric. These mechanisms make the domi- nant contribution to c′(E ) near the top of V1 and bottom of C1 bands, where the Berry cur- vature is peaked (Fig. 4, A and B). The con- tributions from the van Vleck paramagnetism and k-space energy polarization are dominant in the middle of the V1 band, whereas the well- known Landau-Peierls susceptibility is sub- stantial at the band edges. We attribute the large value of c′(E) to the large lattice constant a of the moiré potential. This can be roughly understood with dimen- sional analysis: for a single band model, the total orbital magnetic susceptibility c general- ly scales as Wa2, where W is the bandwidth, so that the derivative of c with respect to energy then scales as a2, which can be large for moiré quantum systems. For a more in-depth analysis, see fig. S10, A and B, which shows that the maximum value of c′ increases dra- matically with increasing moiré wavelength, indicating that future measurements of orbi- tal magnetism should be even more observable at larger wavelengths. MQM holds great opportunities for efficient magnetization control, topological transport phenomena, the quantum anomalous Hall effect, and technological applications in mag- netoelectric, magneto-optics, and topological spintronics. Understanding the underlying band topology and magnetic response functions are the key metrological capability required to practically harness all these opportunities. Our measurements and analysis point a way forward to this goal. RE FERENCES AND NOTES frontiers-in-synthetic-moire-quantum-matter-proceedings- of-a-workshop. 2. D. M. Kennes et al., Nat. Phys. 17, 155–163 (2021). 3. R. Bistritzer, A. H. MacDonald, Proc. Natl. Acad. Sci. U.S.A. 108, 12233–12237 (2011). 4. Y. Cao et al., Nature 556, 80–84 (2018). 5. Y. Cao et al., Nature 556, 43–50 (2018). 6. J. M. Park, Y. Cao, K. Watanabe, T. Taniguchi, P. Jarillo-Herrero, Nature 590, 249–255 (2021). 7. Y. Cao, J. M. Park, K. Watanabe, T. Taniguchi, P. Jarillo-Herrero, Nature 595, 526–531 (2021). 8. H. Zhou, T. Xie, T. Taniguchi, K. Watanabe, A. F. Young, Nature 598, 434–438 (2021). J. M. Park et al., Nat. Mater. 21, 877–883 (2022). 9. 10. G. W. Burg et al., Nat. Mater. 21, 884–889 (2022). 11. Y. Zhang et al., Science 377, 1538–1543 (2022). 12. Y.-H. Zhang, D. Mao, Y. Cao, P. Jarillo-Herrero, T. Senthil, Phys. Rev. B 99, 075127 (2019). 13. M. V. Berry, Proc.- Royal Soc., Math. Phys. Eng. Sci. 392, 45–57 (1984). 14. A. L. Sharpe et al., Science 365, 605–608 (2019). 15. M. Serlin et al., Science 367, 900–903 (2020). 16. S. Sinha et al., Nat. Commun. 11, 5548 (2020). 17. D. Xiao, M.-C. Chang, Q. Niu, Rev. Mod. Phys. 82, 1959–2007 (2010). 18. Y. Gao, S. A. Yang, Q. Niu, Phys. Rev. B Condens. Matter Mater. Phys. 91, 214405 (2015). 19. L. Onsager, Lond. Edinb. Dublin Philos. Mag. J. Sci. 43, 1006–1008 (1952). 20. L. M. Roth, Phys. Rev. 145, 434–448 (1966). 21. Y. Gao, Q. Niu, Proc. Natl. Acad. Sci. U.S.A. 114, 7295–7300 (2017). 22. C. Shen et al., Nat. Phys. 16, 520–525 (2020). 23. X. Liu et al., Nature 583, 221–225 (2020). 24. M. He et al., Nat. Phys. 17, 26–30 (2021). 25. Y. Cao et al., Nature 583, 215–220 (2020). 26. P. C. Adak et al., Phys. Rev. B 101, 125428 (2020). 27. S. Sinha et al., Nat. Phys. 18, 765–770 (2022). 28. C. Zhang et al., Nat. Commun. 12, 2516 (2021). 29. X. Liu et al., Nat. Commun. 12, 2732 (2021). 30. C. Rubio-Verdú et al., Nat. Phys. 18, 196–202 (2021). 31. P. Rickhaus et al., Science 373, 1257–1260 (2021). 32. J. Y. Lee et al., Nat. Commun. 10, 5333 (2019). 33. M. Koshino, Phys. Rev. B 99, 235406 (2019). 34. N. R. Chebrolu, B. L. Chittari, J. Jung, Phys. Rev. B 99, 235417 (2019). 35. Y. W. Choi, H. J. Choi, Phys. Rev. B 100, 201402 (2019). 36. F. J. Culchac, R. R. Del Grande, R. B. Capaz, L. Chico, E. S. Morell, Nanoscale 12, 5014–5020 (2020). 37. F. Haddadi, Q. Wu, A. J. Kruchkov, O. V. Yazyev, Nano Lett. 20, 2410–2415 (2020). 38. J. A. Crosse, N. Nakatsuji, M. Koshino, P. Moon, Phys. Rev. B 102, 035421 (2020). 39. M. Kuiri et al., Nat. Commun. 13, 6468 (2022). 40. Y. J. Song et al., Rev. Sci. Instrum. 81, 121101 (2010). 41. J. Schwenk et al., Rev. Sci. Instrum. 91, 071101 (2020). 42. Additional supplementary text and data are available in the supplementary materials. 43. J. A. Stroscio, W. J. Kaiser, Scanning Tunneling Microscopy: Volume 27 (Methods of Experimental Physics) (Academic Press, 1993). 44. R. Bistritzer, A. H. MacDonald, Phys. Rev. B Condens. Matter Mater. Phys. 84, 035440 (2011). 45. M.-C. Chang, Q. Niu, J. Phys. Condens. Matter 20, 193202 (2008). 46. F. Piéchon, A. Raoux, J.-N. Fuchs, G. Montambaux, Phys. Rev. B 94, 134423 (2016). 47. M. Ogata, H. Fukuyama, J. Phys. Soc. Jpn. 84, 124708 (2015). 48. F. Wilczek, A. Shapere, Geometric Phases in Physics in Advanced Series in Mathematical Physics (World Scientific, 1989), vol. 5; https://www.worldscientific.com/worldscibooks/ 10.1142/0613. 49. H. Fukuyama, Prog. Theor. Phys. 45, 704–729 (1971). 50. J.-N. Fuchs, F. Piéchon, G. Montambaux, SciPost Physics 4, 1. National Academies of Sciences, Engineering, and 024 (2018). Medicine, Theoretical Considerations Concerning Moiré Quantum Materials in Frontiers in Synthetic Moiré Quantum Matter: Proceedings of a Workshop (2022); https://nap.nationalacademies.org/catalog/26594/ 51. M. R. Slot et al., Replication Data for A Quantum Ruler for Orbital Magnetism in Moiré Quantum Matter, George Mason University Dataverse (2023); https://doi.org/10.13021/ orc2020/RLZTDF. Slot et al., Science 382, 81–87 (2023) 6 October 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E ACKN OW LEDG MEN TS We thank F. J. Giessibl for providing the qPlus AFM sensors and general advice on AFM measurements. Funding: University of Maryland and the National Institute of Standards and Technology (NIST) PREP program grant no. 70NANB18H165 (Y.M., E.M.S., and E.S.); Georgetown University and NIST PREP program grant no. 70NANB18H161 (M.R.S.); University of Maryland and NIST Joint Quantum Institute grant no. 70NANB21H126 (S.K. and D.Y.); Office of Naval Research grant no. N00014-20-1-2352 (S.K. and D.Y.); Japan Society for the Promotion of Science KAKENHI grant nos. 19H05790, 20H00354, and 21H05233 (T.T. and K.W.); Dutch Research Council (NWO) through Rubicon grant no. 019.193EN.026 (M.R.S.); University of Reno start-up grant no. PG19012 (Y.B.). Author contributions: Conceptualization: F.G., P.M.H., J.A.S., and N.B.Z.; Methodology: P.M.H. and Y.B.; Investigation: M.R.S., Y.M., S.K., D.T.W., E.S., S.T.L., E.M.S., D.Y., S.R.B., and F.G.; Visualization: Y.M., M.R.S., P.M.H., and J.A.S.; Project administration: J.A.S.; Supervision: J.A.S. and N.B.Z.; Resources: K.W. and T.T.; Software: P.M.H., Y.B., M.R.S., Y.M., S.K., D.T.W., and J.A.S.; Writing – original draft: M.R.S., Y.M., P.M.H., N.B.Z., F.G., and J.A.S.; Writing – review and editing: M.R.S., Y.M., P.M.H., S.K., D.T.W., E.S., S.T.L., E.M.S., D.Y., S.R.B., K.W., T.T., T.B., N.B.Z., F.G., and J.A.S. Competing interests: The authors declare that they have no competing interests. Data and materials availability: The data from this study and the code used to generate theoretical results in this study are available at the George Mason University Dataverse (51). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf2040 Materials and Methods Supplementary Text Figs. S1 to S13 References (52–56) Submitted 5 October 2022; resubmitted 1 March 2023 Accepted 30 August 2023 10.1126/science.adf2040 Slot et al., Science 382, 81–87 (2023) 6 October 2023 7 of 7
10.1126_science.ade4441
RES EARCH 3D PRINTING A silicone-based support material eliminates interfacial instabilities in 3D silicone printing Senthilkumar Duraivel1, Dimitri Laurent2, Didier A. Rajon2, Georg M. Scheutz3, Abhishek M. Shetty4, Brent S. Sumerlin3, Scott A. Banks5, Frank J. Bova2, Thomas E. Angelini1,5,6* Among the diverse areas of 3D printing, high-quality silicone printing is one of the least available and most restrictive. However, silicone-based components are integral to numerous advanced technologies and everyday consumer products. We developed a silicone 3D printing technique that produces precise, accurate, strong, and functional structures made from several commercially available silicone formulations. To achieve this level of performance, we developed a support material made from a silicone oil emulsion. This material exhibits negligible interfacial tension against silicone-based inks, eliminating the disruptive forces that often drive printed silicone features to deform and break apart. The versatility of this approach enables the use of established silicone formulations in fabricating complex structures and features as small as 8 micrometers in diameter. S ilicone elastomer’s resistance to heat, chemical agents, weathering, ozone, moisture, and ultraviolet (UV) irradia- tion makes it critical for manufacturing countless products, including electronic devices, automobiles, aircraft, and medical de- vices (1). Silicone elastomers have been used in medical devices for many years (2), and their applications include embedded sensors (3), flexible electronics (4), soft robotics (5), and additive manufacturing (6). Silicone structures can be fabricated by using conventional tech- niques such as molding, or advanced techniques such as soft lithography and 3D printing (7–9). However, 3D printing with silicone generally results in low-quality products because of chal- lenges created by the interfacial behaviors of silicone pre-elastomer in its liquid state. These challenges can be partially addressed by using an embedding support material that flows around the translating printing nozzles while trapping deposited inks in space, providing stability to printed structures (10–14). How- ever, even under such stabilizing conditions, the interfacial tension between printed inks and their support media drives the deforma- tion and breakup of printed structures before they solidify (Fig. 1, A and B) (9, 15). Modifying silicone inks with additives can stabilize 3D printed structures (16, 17), yet a versatile ap- proach to additive manufacturing with un- modified silicone inks remains elusive. One 1Department of Materials Science and Engineering, University of Florida, Gainesville, FL 32603, USA. 2Department of Neurosurgery, University of Florida College of Medicine, Gainesville, FL 32608, USA. 3George and Josephine Butler Polymer Research Laboratory, Center for Macromolecular Science and Engineering, Department of Chemistry, University of Florida, Gainesville, FL 32611, USA. 4Advanced Technical Center, Anton Paar USA, Ashland, VA 23005, USA. 5Department of Mechanical and Aerospace Engineering, University of Florida, Gainesville, FL 32611, USA. 6J. Crayton Pruitt Family Department of Biomedical Engineering, University of Florida, Gainesville, FL 32611, USA. *Corresponding author. Email: t.e.angelini@ufl.edu route to achieving high-quality 3D silicone printing without ink modification is to elim- inate the disruptive role of interfacial tension by using support materials that are chemically similar to the printed inks they stabilize (Fig. 1C). Thus, there is a critical need to develop support materials that are chemically similar to poly(dimethylsiloxane) (PDMS) inks. We describe a method for 3D printing pre- cise, intricately detailed structures made from PDMS that makes use of a support material ex- hibiting negligible interfacial tension when in contact with silicone inks. We call this method additive manufacturing at ultralow interfacial tension (AMULIT). The AMULIT support mate- rial is a packed inverse emulsion composed of aqueous droplets in a continuum of silicone oil. The ultralow interfacial tension between the AMULIT support material and PDMS inks enabled us to print features with diameters as small as 8 mm. We achieved high-performance printing by tuning the elasticity and flow properties of this support material, which allowed us to fabricate complicated shapes such as brain aneurysm models and func- tional trileaflet heart valves. We demonstrated that the AMULIT technique does not require specialized inks by using several different com- mercially available PDMS formulations to print various structures. With mechanical testing, we found that 3D printed structures produced by using AMULIT were more extensible than their molded counterparts and equally robust. We also found that these structures have a smooth surface finish at the macroscale and microscale roughness, which is facilitated by the low interfacial tension between PDMS inks and the AMULIT support medium. Our results show that the AMULIT 3D printing technique could be used to fabricate intricate silicone structures for biomaterial design and surgical simulators, and they introduce the possibility of expanding the method for printing with other materials. Results Formulation and testing of AMULIT support material To formulate an AMULIT support medium for 3D printing with PDMS inks, we prepared inverse emulsions in which silicone oil was the continuous phase and varied the aqueous drop- let packing fraction, f, and the average drop- let radius, a, between samples; f and a can be tuned independently to determine an emul- sion’s rheological properties and its corre- sponding performance as a printing support medium (18). We expected a to strongly influ- ence the printed feature roughness because the material interfaces will not spontaneously flatten under conditions of ultralow interfacial tension. Thus, we formulated small emulsion droplets and chose f on the basis of the emul- sions’ rheological properties (fig. S1). The elastic shear modulus, G′, and yield stress, sy, of each formulation, were measured with rheological tests (materials and methods and fig. S2). For AMULIT printing, we chose an emulsion having sy = 9 Pa and G′ = 320 Pa; the emulsion with these properties is weak enough to flow around a translating printing needle yet strong enough to support complex 3D printed structures (9, 10). For this formulation, we estimated the Reynolds number near the translating nozzle during a typical print to be 10−6 to 10−2, which indicates that irregular flow patterns should be suppressed (supplementary text). For all formulations, we found that emulsions made from pure water droplets in silicone oil were extremely cloudy and inhibited visualizing the printing process. To make optically clear emul- sions, we matched the refractive indices of the two phases by adding glycerol to the droplets, which allowed the 3D printing process to be imaged at the macroscale with photography and at the microscale with confocal fluores- cence microscopy (CFM) (Fig. 1, D to G, and fig S3). To test the role of interfacial tension in em- bedded 3D printing, we compared the perform- ance of the AMULIT support medium with an all-aqueous support medium made from packed hydrogel microparticles swollen in water. In both cases, we 3D printed features made from a fluorescent PDMS liquid and imaged the ink-support interfaces using CFM (materials and methods). We formulated the packed microgels to have sy = 10 Pa and G′ = 550 Pa, values comparable to those of the AMULIT material. Examining the 3D fluores- cence images, we found that printed silicone features broke up and formed spherical drop- lets within the aqueous support. When a liquid ink is printed into a packed granular support medium, the smallest stable feature has a diameter given by dmin ≈ 2g/sy, where g is the interfacial tension between the ink and the support medium (15). For the aqueous medium, g = 25 mN/m, so dmin was ≈5 mm, Duraivel et al., Science 379, 1248–1252 (2023) 24 March 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Interfacial tension drives feature breakup in embedded 3D printing. (A) High interfacial tension between aqueous support materials and silicone inks destabilizes 3D printed features, driving them to break into spherical droplets. (B) Intermediate interfacial tension between organic support materials and silicone inks provides some stability but limits minimum stable feature size. (C) Ultralow interfacial tension between silicone oil–based support materials and silicone inks eliminates interfacial instabilities, removing the limits on minimum stable feature size. (D) CFM image showing silicone-based inks (green) that break into droplets when printed into support materials made from aqueous microgels (red). (E) A droplet digitally isolated from the support and examined from different angles. The droplet appears nearly spherical and exhibits a smooth surface. (F) By contrast, the silicone-based ink (green) remains continuous and retains its shape indefinitely after printing into a silicone-based support material (red). (G) When viewed from different angles, the printed features exhibit roughness with a characteristic length scale of the microparticles composing the support material, facilitated by ultralow interfacial tension. 50 times the 100-mm diameter of the printed feature (Fig. 1, D and E). Thus, the breakup of the feature into droplets was expected. By con- trast, the 100-mm diameter silicone feature printed into the AMULIT support material re- mained intact, indicating that g < 0.5 mN/m. To better estimate g between a PDMS ink and the AMULIT support medium, we performed a series of test prints in which dmin was mea- sured for multiple values of sy, finding that g ≈ 0.08 mN/m (fig. S4). We also observed that the characteristic roughness length scale at the feature surface was about one order of magnitude smaller than the feature di- ameter, from which we would estimate g ≈ 0.05 mN/m. These results indicate that the AMULIT approach can potentially achieve features 300 to 500 times smaller than those achievable when printing PDMS into an aque- ous support medium having the same material properties. Complex device fabrication using the AMULIT technique The improvement in complexity, quality, and functionality of PDMS vessel models traced in the published literature parallels a decrease in interfacial tension of silicone inks against their embedding materials. For example, hy- drocarbon support materials (9) improved on aqueous support materials (19). As a first test of the AMULIT method’s capabilities, we printed a model brain aneurysm; models with accu- rate vasculature are needed for improved patient simulators to train neurosurgeons in cerebrovascular procedures. Current simulated tissues provide unrealistic tactile feedback, lack small-diameter intracranial angioarchi- tecture, and often exclude the aortic arch and extracranial vascular anatomy that determine which catheters and instruments are used in each procedure (20, 21). To create a model, we collected a 3D an- giogram of a patient’s brain aneurysm using x-ray computed tomography (XRCT). The 3D scan was segmented and processed to create a series of 3D printing trajectories (Fig. 2A and materials and methods). We used Gelest ExSil 100 silicone pre-elastomer, which can be formulated to have material properties that mimic a wide range of tissues. A snapshot from a video of the printing process demon- strates how the translating needle flows easily through the jammed emulsion, which traps the deposited silicone in place (Fig. 2B and movie S1). The printed structure was cured at 60°C for 24 hours and then imaged with XRCT (Fig. 2C). Horizontal and vertical slices through the 3D scan revealed that the highly branched, complex printed network of vessels is hollow, with an average wall thickness of ≈400 mm (Fig. 2D and movie S2). The CT scan of the printed structure was used to create a 3D model for quantitative comparison with the original angiogram. The registration between the patient-derived model and the printed model is excellent; 68% of the printed- surface locations lie within 500 mm of their programmed locations, and 95% lie within 1 mm (Fig. 2, E and F). Our ability to accurately model brain vascu- lature raises the question of whether such fine structures can be manufactured to be both highly compliant and physically robust. The artificial aortic heart valve belongs to a class of devices with such requirements. Native aortic heart valves are subject to dynamic mechanical loads during the cardiac cycle (22). Prosthetic replacement is widely used to treat aortic valve failure, yet the predominantly used mechanical valves and allogeneic- or xenogeneic-tissue valve replacements often result in mechanical failure, hemolysis, blood coagulation, or structural degradation due to calcification. A potential alternative is an arti- ficial silicone valve prosthesis; silicone is es- tablished in vascular applications because of its hemocompatibility and durability (22–27). The AMULIT 3D printing method can be used to replicate the intricate semilunar shape of the thin aortic leaflets in manufactured sili- cone valves. We designed a model heart valve based on physiologically representative dimen- sions of the different valve components (Fig. 2, G and H, and fig. S5) (28). We used a UV- curable silicone formulation, Silopren UV Electro-225-1 (Momentive), as the ink and printed it into the AMULIT material (Fig. 2I). To create highly flexible leaflets, we printed the structure by translating the needle tip at a speed of 2 mm/s and depositing material at a rate of 125 mL/hour, producing features ≈150 mm in diameter. Correspondingly, we chose a layer spacing of 100 mm for good layer adhesion. The printed model was then UV cured, re- moved from the AMULIT material, washed with detergent, and rinsed in deionized water (materials and methods). The cured part had a final wall thickness of ≈250 mm. Despite having very thin, flexible walls, the model valves were physically robust enough to connect to pipe Duraivel et al., Science 379, 1248–1252 (2023) 24 March 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. AMULIT printing of brain aneurysm and aortic heart valve models. (A) Brain aneurysm models for surgical simulations comprise complex, intercon- nected, hollow tubes with intricate details. (B) Photograph of the aneurysm model being printed into the AMULIT material. (C) CT imaging of the 3D printed model within the printing container shows the complexity of the printed aneurysm. (D) Slices through the CT scan show that the printed structure exhibits the hollow channels of the patients’ neurovasculature. (E and F) The printed structure overlays well with the patient’s neurovasculature, and quantitative error analysis demon- strates agreement between the two (±1 mm error range corresponds to 95% of all points). (G and H) A model tricuspid aortic heart valve designed by using the geometric measurements of the native heart valve. (I) A silicone heart valve model printed in a single seamless trajectory with a wall thickness of 250 mm within the AMULIT support medium and cured under a UV lamp. (J and K) Once cured and washed, the valve model is robust enough to be coupled with a water supply, simulating transvalvular flow of the cardiac cycle. The thin leaflets of the valve are observed to open and close during the systolic and diastolic flow of the simulation. explore how well d can be predicted with the AMULIT technique, we printed a series of linear features using the Smooth-On Mold Max 10 PDMS formulation at different com- binations of n and Q and then measured d (Fig. 3A and materials and methods).We pre- dicted the relationship between d, Q, and n, given by p (d/2)2 = Q/n, according to basic fluid continuity. Performing many experi- ments at different combinations of Q and n, we found that this prediction matched the measured feature diameter very well with no adjustable parameters (Fig. 3B). These printed features were stable over time; the change in measured feature size over the course of 120 min postprinting was found to be neg- ligible (fig. S6). We were able to fabricate stable silicone features as small as 8 mm in di- ameter using the AMULIT printing technique; the smallest stable feature diameter we have seen previously demonstrated with unmodified silicone was 40 mm, although smaller unstable features were also reported (9). A feature di- ameter of 10 mm was previously achieved by modifying silicone ink with emulsion droplets (16, 29). To print these very fine features, we formulated an AMULIT support material with an increased yield stress using droplets 1 mm in diameter (fig. S1); the high-magnification images in Fig. 1F indicate that larger droplets would impose interfacial roughness compara- ble to these small feature diameters. We have shown that highly controlled 3D printing with PDMS is possible with the AMULIT Fig. 3. Control of AMULIT printed feature size. (A) (Left) Intensity-inverted images. These images are averaged along the x axis, yielding an intensity profile across each feature. (Right) A Gaussian function is fit to the intensity profile to determine the diameter of the printed line. We measured printed feature diameter with brightfield microscopy, varying the translation speed, n, of the printing nozzle and the ink deposition rate, Q. (B) Feature diameter of the printed silicone is controllable and can be predicted from a fluid continuity equation with no fitting parameters. fittings and simulate transvalvular blood flow through cyclic pumping of water (movie S3). During the negative flow of the pulse rep- resenting the diastolic cycle, the valves re- mained closed with very little deflection on the thin leaflets (Fig. 2J), and during the posi- tive pulse corresponding to the systolic cycle, the leaflets deflected, opening the valve and letting the water flow (Fig. 2K). AMULIT performance: Feature size and print quality The wall thicknesses of the brain vasculature and heart valve models were set by using a combination of feature diameter and layer spacing. The feature diameter, d, for different prints, can be chosen by selecting a combi- nation of nozzle translation speed, n, and ma- terial deposition rate, Q. To systematically Duraivel et al., Science 379, 1248–1252 (2023) 24 March 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Material and surface properties of the AMULIT printed silicone structures. (A and B) Silicone tensile specimens are subject to unidirectional tensile stress and are stretched to failure. (C) Tensile stress-strain curves of the specimens printed with their features oriented parallel and perpendicular to the tensile force show linear stress-strain relationships at low strains and exhibit an elastic modulus of 28 kPa. (D and E) Surface profiles of the printed silicone heart valves exhibit microroughness with an RMS value of 5.5 mm, likely determined by the emulsion droplet radius and the ultralow interfacial tension with the AMULIT support material. technique, and the functionality of the heart valve model suggests that such structures may be sufficiently compliant and durable for use in applications. To test the mechanical perform- ance of printed silicone structures, we fabri- cated tensile specimens using PlatSil-71 RTV (room-temperature vulcanizing) (Polytek) sili- cone formulation following ASTM standard D412 Type C specifications. To test the role of layer-to-layer adhesion in the mechanical in- tegrity of the samples, we printed them with their extruded features oriented in both the longitudinal and lateral directions with respect to the long axis of the specimen geometry. The printed structures were cured at 60°C for 4 hours and then tested by using an Instron 5943 at a loading rate of 500 mm/min (Fig. 4, A and B). The tensile stress-strain data showed that both the lateral and longitudinal print specimens differed negligibly from one anoth- er and had the same elastic modulus of 28 kPa (Fig. 4C). All printed specimens exhibited linear stress-strain relationships at low strain levels and repeatable stress-strain curves at higher strains, failing at strains greater than 1000%. Comparing these results with the per- formance of molded specimens, we found that all the stress-strain curves had the same shape but that printed structures failed at higher strains than molded structures, whereas molded structures exhibited elastic moduli approxi- mately twice those of printed structures. This softening effect could arise from systematic heterogeneities in the printed structures in- herent to the 3D printing process. Addition- ally, we conducted fatigue tests, imposing 105 cycles of ±10% strain, alternately stretching and buckling the samples. Subsequent tensile tests showed that the printed structures ex- hibited less fatigue than did their molded counterparts; the elastic modulus dropped by 18% for the cast samples and 14% for the printed structures (fig. S7). As a final assessment of the quality of struc- tures fabricated with the AMULIT printing technique, we investigated the surface finish of fabricated parts. The ultralow interfacial tension between the silicone and the AMULIT support material was expected to produce micro- rough surfaces on the printed shapes. Using CFM, we imaged a segment of the heart valve model immersed in a rhodamine solution, vis- ualizing and quantifying the surface roughness in 3D. We found the root mean square (RMS) roughness to be 6.54 ± 0.95 mm (mean and standard error, respectively) which is compa- rable to the average diameter of emulsion droplets used in these tests, ≈4 mm. Thus, we expect a smaller roughness with smaller emul- sion droplets such as those used to print very fine features (Fig. 3B). This value is also com- parable to the roughness of PDMS structures printed into support materials that exhibit a high interfacial tension against silicone inks (9), so it may be limited by other factors. In either case, our results demonstrate that elimi- nating disruptive interfacial driving forces with the AMULIT technique enables precise silicone printing without reducing surface quality or mechanical performance of fabricated struc- tures. The added role of emulsion droplet size in surface roughness may enable a printed structure’s optical properties to be tuned while independently controlling its mechanical per- formance through ink composition or feature diameter. can be used to make precise, smooth, strong, and functional devices from commercially available PDMS formulations. The versatility of the AMULIT technique eliminates the need to formulate specialized PDMS inks for 3D applications and broadens the toolbox for re- searchers and industrial manufacturers seek- ing to 3D print PDMS-based devices, while improving on previous silicone printing meth- ods. The AMULIT strategy hinges on formu- lating support materials that are chemically similar to the inks they support—in this case, PDMS inks printed into a continuum of PDMS oil—although the same principle could be used with aqueous polymers. Despite the chemical similarity between the ink and the support medium, we never observed intermixing be- tween the two materials that interfered with printing quality. The very low Reynolds num- ber exhibited during embedded 3D printing with materials such as those we used should facilitate the formation of ink–support inter- faces (30), potentially stabilized by an effective interfacial tension (31) or a form of liquid– liquid phase separation (32), likely influenced by the jammed emulsion phase. Additionally, weak attractive interactions between the emul- sion droplets may help to retain them on their side of the interfaces (33–35). In the near term, we envision the AMULIT method to be useful in 3D printing for a wide range of applications beyond silicone-based devices, given the diver- sity and availability of polymer systems and the simplicity of formulating AMULIT support materials. Conclusions The AMULIT 3D printing method eliminates the disruptive effects of interfacial tension be- tween printed inks and their support mate- rials. Our results show that AMULIT printing REFERENCES AND NOTES 1. H. H. Moretto, M. Schulze, G. Wagner, in Ullmann’s Encyclopedia of Industrial Chemistry (Wiley VCH, 2011), pp. 23–26. 2. A. Rahimi, A. Mashak, Plast. Rubber Compos. 42, 223–230 (2013). 3. X. Ruan et al., Adv. Mater. Technol. 5, 2000171 (2020). Duraivel et al., Science 379, 1248–1252 (2023) 24 March 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E 4. A. Nathan et al., Proc. IEEE 100, 1486–1517 (2012). 5. O. D. Yirmibeşoğlu et al., in 2018 IEEE International Conference on Soft Robotics (RoboSoft), Livorno, Italy, 24 to 28 April 2018 (IEEE, 2018), pp. 295–302. 6. F. Liravi, E. Toyserkani, Addit. Manuf. 24, 232–242 (2018). 7. V. Ozbolat et al., ACS Biomater. Sci. Eng. 4, 682–693 (2018). 8. T. Femmer, A. J. Kuehne, M. Wessling, Lab Chip 14, 2610–2613 (2014). 9. C. S. O’Bryan et al., Sci. Adv. 3, e1602800 (2017). 10. T. Bhattacharjee et al., Sci. Adv. 1, e1500655 (2015). 11. T. Bhattacharjee et al., Soft Matter 14, 1559–1570 (2018). 12. G. Ovarlez, Q. Barral, P. Coussot, Nat. Mater. 9, 115–119 (2010). 13. L. Mohan, C. Pellet, M. Cloitre, R. Bonnecaze, J. Rheol. (N.Y.N.Y.) 57, 1023–1046 (2013). 14. J. R. Seth, L. Mohan, C. Locatelli-Champagne, M. Cloitre, R. T. Bonnecaze, Nat. Mater. 10, 838–843 (2011). 15. C. S. O’Bryan, A. Brady-Miné, C. J. Tessmann, A. M. Spotz, T. E. Angelini, Soft Matter 17, 3886–3894 (2021). 16. B. M. Rauzan, A. Z. Nelson, S. E. Lehman, R. H. Ewoldt, R. G. Nuzzo, Adv. Funct. Mater. 28, 1707032 (2018). 17. C. Perrinet, E. J. Courtial, A. Colly, C. Marquette, R. Fulchiron, Adv. Mater. Technol. 5, 1901080 (2020). 18. T. G. Mason, J. Bibette, D. A. Weitz, Phys. Rev. Lett. 75, 2051–2054 (1995). 19. T. J. Hinton, A. Hudson, K. Pusch, A. Lee, A. W. Feinberg, ACS Biomater. Sci. Eng. 2, 1781–1786 (2016). 20. M. Waqas et al., Neurosurgery 87, E445–E453 (2020). 21. T. Scullen, J. D. Nerva, P. S. Amenta, A. S. Dumont, Neurosurgery 87, E454–E455 (2020). 22. F. B. Coulter et al., Matter 1, 266–279 (2019). 23. M. A. Punchard et al., Ann. Biomed. Eng. 37, 1322–1330 (2009). 24. R. N. Ghriallais, L. McNamara, M. Bruzzi, J. R. Soc. Interface 10, 20120965 (2013). 25. L. Rouleau, J. Rossi, R. L. Leask, J. Biomech. Eng. 132, 071015 (2010). 26. J. Martorell et al., Cardiovasc. Res. 103, 37–46 (2014). 27. M. Balcells et al., Circulation 121, 2192–2199 (2010). 28. W. Sun, C. Martin, T. Pham, Annu. Rev. Biomed. Eng. 16, 53–76 (2014). 29. Z.-T. Xie, D.-H. Kang, M. Matsusaki, Soft Matter 17, 8769–8785 (2021). 30. K. J. LeBlanc et al., ACS Biomater. Sci. Eng. 2, 1796–1799 (2016). 31. D. Truzzolillo, L. Cipelletti, Soft Matter 13, 13–21 (2016). 32. S. Duraivel et al., Biophys. Rev. 3, 031307 (2022). 33. V. V. Erramreddy, S. Ghosh, Langmuir 30, 11062–11074 (2014). 34. A. Z. Nelson, R. H. Ewoldt, Soft Matter 13, 7578–7594 (2017). 35. A. Z. Nelson et al., Curr. Opin. Solid State Mater. Sci. 23, 100758 (2019). 36. S. Duraivel, T. E. Angelini, Heart Valve Maker, v2.0, Zenodo (2023); https://doi.org/10.5281/zenodo.7643835. ACKN OWLED GMEN TS The authors thank Anton Paar for use of their MCR 702 rheometer through the Anton Paar VIP research program. Funding: No funding was received. Author contributions: Conceptualization: S.D. and T.E.A. Methodology: S.D., D.L., D.A.R., G.M.S., A.M.S., B.S.S., S.A.B., F.J.B., and T.E.A. Investigation: S.D., D.L., D.A.R., A.M.S., and T.E.A. Visualization: S.D., D.L., D.A.R., A.M.S., B.S.S., S.A.B., F.J.B., and T.E.A. Funding acquisition: B.S.S., S.A.B., F.J.B., and T.E.A. Project administration: T.E.A. Supervision: B.S.S., F.J.B., and T.E.A. Writing – original draft: S.D. and T.E.A. Writing – review and editing: S.D., D.L., D.A.R., G.M.S., A.M.S., B.S.S., S.A.B., F.J.B., and T.E.A. Competing interests: S.D., B.S.S., and T.E.A are inventors on a US patent application (PCT/US2021/037346). All other authors declare that they have no competing interests. Data and materials availability: All data are available in the manuscript and the supplementary material; code for the design and generation of print trajectories for the heart valve model is freely accessible at Zenodo (36). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/ science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.ade4441 Materials and Methods Supplementary Text Figs. S1 to S7 Movies S1 to S3 Submitted 17 August 2022; accepted 17 February 2023 10.1126/science.ade4441 Duraivel et al., Science 379, 1248–1252 (2023) 24 March 2023 5 of 5
10.1126_science.adf3971
RES EARCH R E S E A R C H A R T I C L E S U M M A R Y ◥ CELL BIOLOGY Protein import into peroxisomes occurs through a nuclear pore–like phase Yuan Gao†, Michael L. Skowyra†, Peiqiang Feng, Tom A. Rapoport* INTRODUCTION: Peroxisomes are organelles that are enclosed by a single membrane and exist in nearly all eukaryotic cells. They per- form important functions related to lipid metabolism and redox homeostasis, among others. Peroxisomes are also vital for human health; inborn defects in peroxisome biogene- sis cause devastating and often life-threatening disorders such as the Zellweger spectrum. Per- oxisomal enzymes are made in the cytosol, where they are recognized by PEX5 and re- lated mobile receptors and are shuttled by the receptors into the organelle. How the re- ceptors move proteins across the peroxi- somal membrane has been a long-standing question, particularly because peroxisomes can mysteriously import folded or even oligo- meric proteins. This property fundamentally differs from the way in which proteins are imported into the endoplasmic reticulum or mitochondria. RATIONALE: Protein translocation into perox- isomes requires a number of components embedded in the peroxisomal membrane— notably PEX13. We recognized that PEX13 contains an extensive unstructured region that is enriched in the amino acids tyrosine (Y) and glycine (G). This tyrosine- and glycine-rich YG domain resembles the phenylalanine (F)– and glycine-rich FG domains of nucleoporins that reside in nuclear pores. FG domains form a meshwork inside nuclear pores that restricts passage of soluble molecules yet allows nu- clear transport receptors to diffuse through and bring cargo along. By analogy to FG do- mains, we hypothesized that the YG domain of PEX13 might form a similarly selective phase on peroxisomes, through which import recep- tors could move folded proteins across the peroxisomal membrane. RESULTS: We show that the YG domain is found in PEX13 proteins from all classes of eukaryotic organisms. Our analysis reveals that this domain is distinguished by the presence of repeated aromatic amino acids, predomi- nantly tyrosines, which are separated from one another by short flexible linkers composed mostly of glycines and serines. The tyrosine residues are essential for peroxisomal protein import in the yeast Saccharomyces cerevisiae. YG domains from different organisms vary extensively in their amino acid sequence and in the number of tyrosines, yet despite this diversity, the domains are functionally inter- changeable in yeast. The YG domain is followed by a long amphipathic helix that we show is PEX5 Cargo PEX13 NTR Cargo Nuclear pore Peroxisome Nucleus Peroxisomal protein import resembles nuclear transport. A dense meshwork is formed in the peroxisome’s membrane by the YG domain of multiple copies of the peroxisomal protein PEX13. This meshwork functions as a conduit through which the import receptor PEX5 can selectively diffuse and deliver bound cargo into the peroxisome. The process is similar to how a nuclear transport receptor (NTR) moves through the FG meshwork inside a nuclear pore. also conserved across eukaryotic organisms and required for peroxisomal import. Using disulfide-mediated cross-linking, we demonstrate that YG domains from multiple PEX13 molecules associate with one another in the membrane of yeast peroxisomes. This interaction leads to the formation of a dense meshwork, which excludes large soluble mate- rial such as proteases and bulky molecules but remains permeable to small molecules that are less than 2 kD in size. The formation of the meshwork and its barrier properties depend on the aromatic residues within the YG do- main. Using protease-protection and cysteine- modification experiments, we further show that PEX13 resides in the peroxisomal mem- brane in two opposite orientations, an unusual feature among membrane proteins. PEX13 molecules of both orientations associate with one another, an interaction that involves the proteins’ YG domains. Crucially, we find that the purified YG do- main forms a hydrogel that is held together by the domain’s tyrosine residues. The import receptor PEX5 fused to green fluorescent pro- tein (GFP) can selectively partition into the hydrogel, whereas GFP alone is excluded. The entry of PEX5 into the hydrogel requires con- served aromatic WXXXF/Y motifs (where W is tryptophan and X denotes any residue) in the receptor’s flexible N-terminal region. Further- more, PEX5 can also bring cargo proteins into the YG hydrogel. Delivery of cargo into the gel requires the presence of WXXXF/Y motifs in the receptor. CONCLUSION: Our results with intact yeast mem- branes and synthetic hydrogels reveal that per- oxisomal import resembles transport through the nuclear pore. A dense meshwork—a selec- tive phase—is suspended within the peroxi- somal membrane by multiple PEX13 molecules of opposite orientations. This meshwork pro- vides an aqueous conduit across the membrane into the peroxisomal lumen. The walls of the proposed conduit might be assembled from the protein’s conserved amphipathic helix. Peroxi- somal import receptors such as PEX5 can dif- fuse through the meshwork and bring cargo along, using their WXXXF/Y motifs to locally dissolve the cohesive interactions holding the meshwork together. This mechanism ex- plains how folded and oligomeric proteins are imported into peroxisomes and represents a previously unidentified principle by which proteins cross membranes.▪ The list of author affiliations is available in the full article online. *Corresponding author. Email: tom_rapoport@hms.harvard.edu †These authors contributed equally to this work. Cite this article as Y. Gao et al., Science 378, eadf3971 (2022). DOI: 10.1126/science.adf3971 READ THE FULL ARTICLE AT https://doi.org/10.1126/science.adf3971 Gao et al., Science 378, 1187 (2022) 16 December 2022 1 of 1 RES EARCH R E S E A R C H A R T I C L E ◥ CELL BIOLOGY Protein import into peroxisomes occurs through a nuclear pore–like phase Yuan Gao1,2†, Michael L. Skowyra1,2†, Peiqiang Feng1,2, Tom A. Rapoport1,2* Peroxisomes are ubiquitous organelles whose dysfunction causes fatal human diseases. Most peroxisomal proteins are imported from the cytosol in a folded state by the soluble receptor PEX5. How folded cargo crosses the membrane is unknown. Here, we show that peroxisomal import is similar to nuclear transport. The peroxisomal membrane protein PEX13 contains a conserved tyrosine (Y)– and glycine (G)–rich YG domain, which forms a selective phase resembling that formed by phenylalanine- glycine (FG) repeats within nuclear pores. PEX13 resides in the membrane in two orientations that oligomerize and suspend the YG meshwork within the lipid bilayer. Purified YG domains form hydrogels into which PEX5 selectively partitions, by using conserved aromatic amino acid motifs, bringing cargo along. The YG meshwork thus forms an aqueous conduit through which PEX5 delivers folded proteins into peroxisomes. P eroxisomes are organelles enclosed by a single membrane and are found in most eukaryotic cells (1). They provide vital functions, including fatty acid oxi- dation (2) and detoxification of reactive oxygen species (3). Peroxisomes are essential for human health: Various debilitating and often fatal disorders, notably the Zellweger spectrum, arise from the defective import of enzymes into the peroxisomal lumen, other- wise known as the matrix (4). Matrix proteins are made in the cytosol and then imported into peroxisomes. Most contain a type 1 peroxisome targeting signal (PTS1) at their C terminus, which comprises the amino acid sequence Ser-Lys-Leu (SKL) or variants of it (5). The PTS1 is recognized in the cytosol by the soluble receptor PEX5 through the recep- tor’s tetratricopeptide repeat (TPR) domain (6). PEX5 also contains a flexible N-terminal region that includes several aromatic motifs conforming to the amino acid sequence WXXXF/Y (where W is Trp, X denotes any residue, F is Phe, and Y is Tyr) (7). Some matrix proteins contain an alternative N-terminal signal called PTS2, whose recognition requires the adapter PEX7 (8). In humans, PEX7 binds a short motif in PEX5, whereas in many fungi, PEX7 associates with PEX5 paralogs that lack a TPR domain but retain the unstructured N-terminal region with its characteristic WXXXF/Y motifs (9). PEX5 is recruited to peroxisomes by the membrane proteins PEX13 and PEX14 (and PEX17 in yeast) (7). Recruitment requires the 1Department of Cell Biology, Harvard Medical School, Boston, MA 02115, USA. 2Howard Hughes Medical Institute, Harvard Medical School, Boston, MA 02115, USA. *Corresponding author. Email: tom_rapoport@hms.harvard.edu †These authors contributed equally to this work. receptor’s WXXXF/Y motifs (10) and is fol- lowed by translocation of the cargo-bound receptor completely into the matrix (10–12). To return to the cytosol, PEX5 is ubiquitinated by the PEX2-PEX10-PEX12 ubiquitin ligase com- plex (13) and pulled out through a pore in the ligase complex (14) by a hexameric adenosine triphosphatase (ATPase) consisting of alternat- ing copies of PEX1 and PEX6 (15). Deubiquiti- nation in the cytosol resets PEX5 for a new import cycle (16, 17). How PEX5 crosses the peroxisomal mem- brane to deliver cargo into the lumen has been a long-standing question. Particularly myste- rious is the receptor’s ability to import folded or oligomeric proteins (18), or even gold beads (19). Thus, translocation into peroxisomes fun- damentally differs from that into the endoplas- mic reticulum (ER) or mitochondria, which can only import proteins in an unfolded con- formation (20, 21). It is also puzzling that trans- location across the peroxisomal membrane does not require nucleotide hydrolysis (22), even though import occurs against a concen- tration gradient of the cargo (23). Translocation into peroxisomes is thought to be mediated by PEX13 or PEX14. Although PEX14 has historically been favored (24), the protein lacks obvious features that could form an aqueous conduit for moving hydrophilic proteins across the membrane (25). In addi- tion, PEX14 may be dispensable for import in some organisms (26, 27). PEX13, by contrast, is essential for import in all organisms that have been tested (28, 29). Curiously, PEX13 contains an unstructured N-terminal region of unknown function that is required for import (30) and is enriched in the amino acids tyrosine and glycine (31). We noted that this tyrosine- and glycine- rich YG domain resembles the phenylalanine- and glycine-rich FG domains of nucleoporins, which reside within the nuclear pore and form a meshwork that restricts the entry of large molecules into the nucleus (32). This mesh- work is locally broken by nuclear transport receptors (NTRs), allowing them to rapidly dif- fuse through nuclear pores along with bound cargo (33). Here, we show that the YG domain of PEX13 forms a similarly selective phase on peroxi- somes. Our results lead to a model whereby the YG phase is locally disrupted by aromatic residues in the receptor’s WXXXF/Y motifs, allowing PEX5 to diffuse across the membrane into the matrix and carry cargo along. This mechanism explains how folded and oligomeric proteins are imported into peroxisomes. The YG domain of PEX13 is conserved and essential for peroxisomal import The YG domain is found in PEX13 homologs from species representative of all eukaryotic clades (Fig. 1A). The domain is characterized by a preponderance of aromatic amino acids, predominantly tyrosine, but phenylalanines also occur in some organisms (Fig. 1A and fig. S1A). The number and position of the aromatic residues vary (Fig. 1A). The aro- matic residues are separated by short linkers of about four amino acids (fig. S1B), which are enriched in small residues, notably glycine and serine, and lack charged residues (Fig. 1A and fig. S1A). These properties are generally similar to those of cohesive nucleoporin FG- repeat domains (fig. S1C), except that the linkers between FG repeats are longer (fig. S1D). The YG domain is followed by a long am- phipathic helix (AH) (Fig. 1A and fig. S2), which has clearly defined hydrophobic and hydrophilic surfaces (fig. S3). Whether the AH consists of a straight helical segment or has a kink in the middle is unclear (fig. S3). Downstream of the AH, most PEX13 homologs contain a canoni- cal transmembrane segment (TM) and an SH3 domain that binds PEX5 (28) (Fig. 1A and fig. S2), although these are absent in some plants (34) (fig. S2). Thus, only the YG domain and the AH are strictly conserved. The YG domain is necessary for peroxi- somal import, which we measured in the yeast Saccharomyces cerevisiae using an engineered pathway that generates a pigment when import is impaired (35). When all 14 tyrosines in the YG domain were mutated to serines, import was abolished (Fig. 1B), whereas converting all of the tyrosines into phenylalanines caused only a modest defect (Fig. 1B). Thus, the residues’ aromaticity is critical. A minimum of 11 tyrosines seems to be required in yeast (fig. S4A), but their position can be varied. The YG domain could also be replaced by the analogous domain from other organisms (Fig. 1C), albeit with variable ef- ficiency, indicating that the domain’s function is conserved despite its sequence diversity. Gao et al., Science 378, eadf3971 (2022) 16 December 2022 1 of 14 RES EARCH | R E S E A R C H A R T I C L E A unstructured PEX13 N YG AH TM SH3 S. cerevisiae (yeast) 71 141 214 263 285 303 374 yeast Torula. Pichia amoeba algae plant fruit fly human B YG AH TM SH3 * * * * * * pex13 WT Y S Y F AH mut AH mut SH3 N Import activity (%) 50 100 0 C Import activity (%) 50 100 0 pex13 yeast human fruit fly Torula. plant algae Pichia amoeba YG Fig. 1. The YG domain of PEX13 is required for peroxisomal protein import. (A) The domain organization of PEX13 from S. cerevisiae (yeast) is shown at the top. Numbers denote amino acid coordinates. Sequences of the YG domain in PEX13 homologs from organisms representing the indicated eukaryotic clades (Torula., Torulaspora) are shown at the bottom. (B) Peroxisomal protein import activity (mean ± SE of three experiments) in yeast cells expressing the indicated PEX13 mutants compared with wild-type (WT) cells and a PEX13 knockout strain (pex13D). Y→S and Y→F denote conversion of all tyrosines (positions indicated by vertical bars) in the YG domain into serines or phenylalanines, respectively. AH mut z and ϕ denote conversion of four hydrophilic residues in the AH into alanine or two hydrophobic residues into glutamate, respectively. (C) Import activity in yeast cells expressing PEX13 chimeras with YG domains from the indicated organisms. Single-letter abbreviations for the amino acid residues are as follows: A, Ala; C, Cys; D, Asp; E, Glu; F, Phe; G, Gly; H, His; I, Ile; K, Lys; L, Leu; M, Met; N, Asn; P, Pro; Q, Gln; R, Arg; S, Ser; T, Thr; V, Val; W, Trp; and Y, Tyr. Taken together, these data demonstrate that the YG domain is a universal and essential feature of PEX13. Import was also abolished by point muta- tions in either the hydrophilic face (mut z) or the hydrophobic face (mut ϕ) of the AH (Fig. 1B), revealing that the AH is likewise essential. Whereas loss of the SH3 domain (DSH3) re- duced import (Fig. 1B) as reported previously (36), deleting the flexible region that precedes the YG domain (DN) had no effect (Fig. 1B), indicating that this region is dispensable. We confirmed that all mutants were similarly expressed (fig. S4, A to C). YG domains of multiple PEX13 molecules interact on peroxisomes To determine whether YG domains form a nuclear pore–like phase on peroxisomes, we first tested whether these domains associate with one another in the peroxisomal mem- brane. Individual cysteines were introduced at either of two positions (residue 104 or 131) in the YG domain of FLAG-tagged yeast PEX13 (which lacks natural cysteines) (Fig. 2A), without affecting import activity (fig. S5A). Formation of disulfide-linked dimers was then assessed using an oxidizing agent (Fig. 2A). Both positions indeed yielded efficient dimerization (>70%) (Fig. 2B, top). Dimers were disulfide- linked because the corresponding bands dis- appeared upon reduction with dithiothreitol (DTT) (Fig. 2B, bottom). Dimerization was also sensitive to detergent (Fig. 2B, lanes 4 and 8), suggesting that the interaction requires an intact membrane. These results thus show that the YG domains of individual PEX13 molecules interact in the peroxisomal membrane. The interaction between YG domains is indeed multivalent and does not involve just two molecules. When two cysteines were intro- duced into the YG domain (at positions 104 and 131), a ladder of disulfide-linked bands was observed (Fig. 2C, right). Some cross-linking occurred even without oxidant (lane 4), sug- gesting that the interaction is highly favored. The largest cross-linked species reveal that more than eight PEX13 molecules interact through their YG domains. Cross-linking was greatly reduced after converting all tyrosines in the YG domain into serines (Fig. 2D, compare lanes 3 and 6 and corresponding quantifica- tion). Cross-linking was unaffected by the ab- sence of other import components, including PEX2, PEX5, PEX14, or PEX17 (fig. S4D), sug- gesting that PEX13 forms oligomers by itself. This conclusion agrees with previous studies showing that PEX13 forms oligomers inde- pendently of other import components (37, 38). Our analysis demonstrates that multiple PEX13 molecules associate with one another on peroxi- somes through their YG domains, analogously to how nucleoporin FG domains interact within the nuclear pore. The YG domain is inaccessible to large molecules We next asked whether the interaction between YG domains restricts access to soluble material, similar to the FG meshwork inside nuclear pores that excludes large proteins and other mole- cules. A 3C protease–cleavage site was inserted at different positions in the YG domain of FLAG-tagged PEX13 or outside this domain near the N terminus (Fig. 3A, scheme). The resulting constructs replaced endogenous PEX13 in yeast and were fully active (fig. S5B). When membranes isolated from the corresponding Gao et al., Science 378, eadf3971 (2022) 16 December 2022 2 of 14 RES EARCH | R E S E A R C H A R T I C L E A B 104 131 FLAG YG AH TM SH3 oxidant quench IP FLAG blot FLAG C cysteine 104 + 131 + DTT – DTT cysteine 104 131 oxidant det det oxidant 90 – DTT 50 90 + DTT 50 2× 1× 1× 260 160 90 50 D 1 WT Y S fraction crosslinked 0 1 2 3 4 5 6 cysteine 104 + 131 WT Y S oxidant 260 160 90 50 50 8×> 5-8× 4× 3× 2× 1× – DTT + DTT 1 2 3 4 5 FLAG 6 7 8 1 2 3 4 FLAG 5 6 1 2 4 3 FLAG 5 6 Fig. 2. The YG domain of PEX13 forms oligomers in the peroxisomal mem- brane. (A) Scheme for the experiments shown in (B) to (D). To test whether YG domains associate with one another, cysteines were introduced at the indicated positions in the YG domain of FLAG-tagged yeast PEX13 (which lacks native cysteines), and the resulting constructs were expressed in yeast. Intact membranes from the corresponding strains were treated with Aldrithiol-4 (oxidant), unreacted cysteines were then quenched with NEM, and the proteins were immunoprecipitated (IP) and analyzed by immunoblotting (blot). (B) As described in (A), using PEX13 proteins containing individual cysteines at the indicated positions. Reactions were performed in the presence of 0, 50, or 200 mM oxidant with or without detergent (det) and resolved by nonreducing (–DTT) or reducing (+DTT) SDS–polyacrylamide gel electrophoresis (SDS-PAGE) before immunoblotting. Monomers (1×) and disulfide-linked dimers (2×) are indicated on the right. Numbers along the left side specify relative molecular weights (in kD). (C) Same as in (B), but with PEX13 containing two cysteines. The number of disulfide-linked molecules is indicated on the right (1× to 8×). (D) Same as in (C), but comparing PEX13 with tyrosines (WT) or serines (Y→S) in the YG domain. Cross-linking efficiency is plotted at the top (mean ± SE of three experiments). strains were treated with the protease, the constructs with N-terminal cleavage sites were readily digested (Fig. 3A, lanes 2 and 6), consistent with the reported cytosolic accessi- bility of the N terminus (39). By contrast, sites in the YG domain were much more resistant to cleavage (lanes 10 and 14) unless detergent was added (lanes 11 and 15). No cleavage oc- curred at any site when the protease was preinactivated by N-ethylmaleimide (NEM; lanes marked by skull and crossbones). These data thus show that the YG domain is not easily accessed by the 30-kD 3C protease. To confirm this result, we probed the accessibility of the YG domain to cysteine- reactive polyethylene glycol (PEGmal). Indi- vidual cysteines were introduced throughout FLAG-tagged PEX13 in yeast (Fig. 3B, scheme), and membranes from the corresponding strains were treated with different sizes of PEGmal. Cysteines near the N terminus were readily modified by all sizes of PEGmal (Fig. 3B, upper two blots) as expected. By contrast, cysteines in the YG domain were barely modified by the two largest PEGmal reagents (5 and 10 kD) unless detergent was included (lanes 1 to 6 in the two middle blots). This resistance to modification depended on size, because the two smallest forms of PEGmal (0.8 and 2 kD) efficiently modified either cysteine (lanes 7 to 12 in the two middle blots). As a control, we tested the accessibility of a cysteine in the TM, which was not modified by PEGmal of any size (bottom blot). Converting all tyrosines in the YG domain to serines (Y→S) increased accessibility to 10-kD PEGmal (Fig. 3C), consistent with this mutant’s reduced self-association (Fig. 2D). Converting all tyrosines to phenylalanines (Y→F) instead improved the resistance to modification (Fig. 3C), again highlighting the importance of the aromatic residues in cement- ing the interaction between PEX13 molecules. The absence of PEX5 (pex5D) had no effect on accessibility (Fig. 3C), suggesting that active import may not be necessary for the YG domain’s exclusion properties. Taken together, our results reveal that multiple YG domains form a meshwork in the peroxisomal mem- brane, which excludes large molecules but not small ones, similarly to the FG meshwork in nuclear pores. To confirm this conclusion, we introduced single cysteines on either side of the trans- membrane segment of the peroxisomal mem- brane protein PEX14 (which also lacks native cysteines) and examined their accessibility to different sizes of PEGmal as above (fig. S6A). The C terminus of PEX14 faces the cytosol (10, 39), and accordingly, a cysteine located near the C terminus (position 242) was readily modified by all sizes of PEGmal in the absence of detergent (fig. S6B, upper blot). By contrast, a cysteine located near the N terminus (posi- tion 65), which is oriented toward the lumen (10, 39), was modified only by the smallest PEGmal tested (fig. S6B, lower blot). In the presence of detergent, both cysteines were indiscriminately modified by all sizes of the reagent. These results thus confirm that the peroxisomal membrane is permeable to small molecules, as suggested before (35, 40), and are Gao et al., Science 378, eadf3971 (2022) 16 December 2022 3 of 14 RES EARCH | R E S E A R C H A R T I C L E A protease site protease 55 36 34 67 96 133 YG AH TM SH3 FLAG protease quench IP FLAG blot FLAG B 39 68 104 131 FLAG YG AH 270 TM SH3 detergent (kD) PEGmal + 10 10 – + 5 5 + 2 + – 0.8 0.8 2 34 det + + + 67 det + + + 96 det + + + 133 det + + + un- cleaved 1 2 3 4 5 6 7 9 8 FLAG 10 11 12 13 14 15 16 C cysteine at position 104 WT Y S Y F pex5 + WT detergent (10-kD) PEGmal + + + + + + + + + + + + 70 55 1 2 3 4 5 7 6 FLAG 8 9 10 11 12 1 0 fraction modified Y S T W Y F pex5 70 55 70 55 70 55 70 55 70 55 39 68 104 cysteine position 131 270 1 2 3 4 5 7 6 FLAG 8 9 10 11 12 Fig. 3. The YG domain is poorly accessible to large molecules. (A) To test accessibility of the YG domain to proteins, a 3C protease–cleavage site was introduced into FLAG-tagged yeast PEX13 at the indicated positions, and the resulting constructs were expressed in yeast. Intact membranes from each strain were first treated with the protease with or without detergent (det) and then quenched with NEM to inactivate the protease. Where indicated (skull and crossbones), NEM was added before the protease. Scissors designate the cleaved forms. (B) Same as in (A), except that accessibility of the YG domain to differently sized molecules was assessed by introducing individual cysteines into FLAG-tagged PEX13 at the positions shown. Membranes from the corresponding strains were treated with different sizes (in kD) of cysteine-reactive PEGmal and then quenched with excess cysteine. Covalent modification of the proteins was visualized by immunoblotting. Modified and unmodified forms of the protein are designated by yellow and white triangles, respectively, in the top blot. (C) Same as in (B), but with PEX13 containing a cysteine at position 104 and tyrosines in the YG domain (WT), or tyrosines mutated to serines (Y→S) or phenylalanines (Y→F). Where indicated, membranes were isolated from a strain lacking PEX5 (pex5D). Modification was performed with 10-kD PEGmal. Modification efficiency is plotted on the right (mean ± SE of four experiments). consistent with a dense meshwork of limited porosity residing in the membrane. PEX13 adopts two orientations in the peroxisomal membrane The YG meshwork must be suspended within the peroxisomal membrane to form a conduit for cargo. To determine how the YG meshwork is formed, we examined the membrane topology of PEX13. The N terminus of the protein was fused to a 3C protease–cleavage site preceded by a streptavidin-binding peptide (SBP) tag, and the C terminus was fused to a tobacco etch virus (TEV) protease–cleavage site followed by a FLAG tag (Fig. 4A, upper scheme). The construct was expressed from the endogenous locus in yeast and supported peroxisomal protein import (fig. S5B). We first ascertained the orientation of the C terminus. Membranes were treated with TEV protease, and PEX13 was then immunopreci- pitated by the N-terminal SBP tag (Fig. 4A). Interestingly, only half of the PEX13 popula- tion was cleaved (lane 2; quantification in fig. S7A). The resistant pool was protected by the membrane, because it could be cleaved in the presence of detergent (Fig. 4A, lane 3). No cleavage occurred when the protease was preinactivated by NEM (lane 4). These data thus suggest that PEX13 resides in the mem- brane in two orientations: one whose C terminus faces the cytosol and another whose C terminus faces the lumen (Fig. 4A, bottom scheme). The interaction between YG domains in the mem- brane might thus be mediated by PEX13 mole- cules of both orientations (see below). To infer the orientation of the N terminus, membranes were treated with 3C protease, and PEX13 was then immunoprecipitated by the C-terminal FLAG tag (Fig. 4A). In contrast to the C terminus, more than 50% of the molecules had their N termini accessible to the protease (lane 6), and the exact propor- tion varied between experiments (fig. S7A). The remaining pool was again cleaved in detergent (Fig. 4A, lane 7). The N terminus might thus not be fixed in one specific orien- tation but may instead be free to move between the cytosol and the lumen (Fig. 4A, bottom scheme). Notably, these results agree with the cytosolic accessibility of the N terminus that was reported above (Fig. 3, A and B). Controls were performed with PEX14 (Fig. 4B) and PEX17 (fig. S7B), which are conven- tional single-pass membrane proteins whose N termini face the lumen and C termini face the cytosol (10, 25, 39). Using analogously tagged constructs that supported peroxisomal import (fig. S5B), we found that the C termini of both proteins were fully cleaved in the ab- sence of detergent, whereas their N termini Gao et al., Science 378, eadf3971 (2022) 16 December 2022 4 of 14 RES EARCH | R E S E A R C H A R T I C L E A 3C N (SBP) PEX13 TEV B 3C PEX14 YG AH TM SH3 C (FLAG) N (SBP) PBD TM CC TEV C (FLAG) protease quench IP FLAG / SBP blot FLAG / SBP IP : SBP IP : FLAG det det IP : SBP IP : FLAG det det N C cytosol N M T or T M lumen N FLAG SBP 55 55 cytosol FLAG lumen SBP 1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8 PEX13 TEV 3C C N D PEX13 229 246 295 55 55 C FLAG YG AH TM SH3 FLAG YG AH TM SH3 det det det cysteine detergent (5-kD) PEGmal 229 + + + 246 + + + 295 + + + 246 + 229 + + + 246 + 295 + + + 55 36 full-length fragment (3C) fragment (TEV) 70 50 C M T N +2 +1 1 2 3 4 5 FLAG 6 7 8 1 2 3 4 5 6 7 8 10 11 12 13 14 15 9 FLAG Fig. 4. PEX13 adopts two transmembrane orientations. (A) Membrane topology determined by protease protection. The indicated protease-cleavage sites (scissors) and epitope tags were introduced into yeast PEX13. Membranes containing this protein were treated with protease with or without detergent (det), the reactions were quenched with NEM, and PEX13 was immunoprecipi- tated (IP); cleavage was visualized by immunoblotting (see scheme). Where indicated (skull and crossbones), NEM was added before the proteases. The position of the C terminus is deduced from cleavage with TEV protease and immunoprecipitation by the N-terminal SBP tag, and the position of the N terminus is deduced from cleavage with 3C protease and immunoprecipitation by the C-terminal FLAG tag. The two inferred orientations of PEX13 are depicted on the right; the N terminus can face either side, whereas the C terminus is fixed in one of two orientations. (B) Same as in (A), but for PEX14. PBD, PEX5- binding domain; TM, transmembrane segment; CC, coiled-coil oligomerization domain. (C) Same as in (A), but with protease sites flanking the TM and a FLAG tag as shown. (D) TM orientation determined by modification of flanking cysteines with membrane-impermeable PEGmal. One or two cysteines were introduced into FLAG-tagged PEX13, as indicated. Membranes were treated with 5-kD PEGmal and then quenched with excess cysteine; single (+1) or double (+2) modification was visualized by immunoblotting. were completely protected unless detergent was included (Fig. 4B and fig. S7B). The topol- ogies of PEX14 and PEX17 thus agree with previous reports. The dual topology of PEX13 is supported by two observed orientations of the protein’s TM. We incorporated 3C protease– and TEV protease– cleavage sites on either side of the TM and added a FLAG tag to the N terminus for im- munodetection (Fig. 4C, scheme). The resulting construct allowed peroxisomal import (fig. S5B). Again, half of the molecules were accessible to either protease in the absence of detergent (Fig. 4C, lanes 2 and 4), whereas the entire population was cleaved when detergent was included (lanes 3 and 5). Notably, addition of both proteases produced both cleavage products (lanes 6 to 8), consistent with two orientations of the protein in the membrane. We confirmed the dual topology of PEX13 by PEGmal accessibility. Single cysteines were introduced on either side of the TM (Fig. 4D, scheme) without affecting import (fig. S5A), and membranes from the corresponding strains were treated with 5-kD PEGmal. Regardless of whether the cysteine resided upstream (position 229 or 246) or downstream (position 295) of the TM, about half of the molecules showed the increase in molecular weight that was expected from cysteine modification (Fig. 4D, left blot). All molecules were modified in detergent. With two cysteines on the same side of the TM (positions 229 and 246), half of the PEX13 molecules were modified on both cysteines, whereas the other half were not modified at all (Fig. 4D, right blot), as expected. By con- trast, with two cysteines on opposite sides (po- sitions 246 and 295), essentially all PEX13 molecules were modified but only on one or the other cysteine. These results confirm that the TM of PEX13 spans the membrane in two pos- sible orientations. Gao et al., Science 378, eadf3971 (2022) 16 December 2022 5 of 14 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. The two orientations of PEX13 are bridged by the YG domain. (A) To test whether the two orientations of PEX13 associate with each other, a single cysteine and a 3C protease–cleavage site were incorporated into FLAG-tagged yeast PEX13 as shown. An SBP tag was included at the C terminus to enhance the size shift after proteolytic cleavage. The resulting construct was integrated into yeast. Intact membranes from the corresponding strain were treated with the protease to reveal the protein’s two orientations, oxidized with Aldrithiol-4 to induce disulfide-bridge formation, and quenched to inactivate the protease and oxidant. (B) Disulfide-linked dimers were visualized by reducing (+DTT) or non- reducing (–DTT) SDS-PAGE and immu- noblotting for the FLAG tag. Where indicated, detergent (det) was included during protease cleavage. Cleaved (scissors) and uncleaved (∅) species are marked on the right. The topologies of the three observed dimers (numbered) are depicted below. A FLAG 104 YG PEX13 3C AH TM SH3 SBP protease oxidize quench IP FLAG B cysteine 104 + oxidant + DTT – DTT det det + + + 1 2 3 1 2 3 4 5 6 1 2 3 FLAG 100 55 cytosol lumen The dual topology does not require forma- tion of the YG meshwork, because both orien- tations were retained after mutating all tyrosines in the YG domain into serines (fig. S8A). Both orientations were also retained in the absence of PEX2, PEX5, PEX8, PEX14, or PEX17 (fig. S8B), indicating that other import components are dispensable. By contrast, in the absence of PEX3, all PEX13 molecules became accessible to the proteases and the abundance of PEX13 was also reduced (fig. S8C). This observation agrees with previous reports of PEX13 being destabilized in the absence of PEX3 (41, 42) and with the proposed role of PEX3 in peroxi- somal membrane protein insertion (43). The YG domain bridges both orientations of PEX13 We next asked whether the two orientations of PEX13 associate with each other, allowing the YG domains of these molecules to interact within the membrane. We introduced a single cysteine into the YG domain and tested whether disulfide-linked dimers would form between the two orientations. To discern both orientations, the C terminus was fused to a 3C protease–cleavage site followed by an SBP tag (to enhance the size shift), and an N-terminal FLAG tag was added for immunodetection (Fig. 5A). Cleavage by the protease produced two bands of equal intensity in a reducing gel (Fig. 5B, left blot), corresponding to the two orientations of the C terminus. Under oxidizing conditions, however, three disulfide-linked bands were seen after cleavage (right blot), corres- ponding to PEX13 dimers in which the two molecules have the same or opposite topologies (Fig. 5B, scheme). These results establish that both orientations of PEX13 directly interact with each other through their YG domains and thereby explain how the YG meshwork is localized to the plane of the membrane. YG domains form hydrogels Given that the nucleoporin FG meshwork can be reconstituted as a hydrogel (44) whose per- meation properties mimic transport through the nuclear pore (45), we wondered whether the YG domain of PEX13 might behave sim- ilarly. After screening expression of YG domains from different organisms, we found that the entire unstructured N-terminal region (includ- ing the YG domain) from Arabidopsis thaliana PEX13 (Fig. 6A) expressed well in bacteria and could be purified under denaturing conditions using a His tag. Notably, the YG domain from this species could complement the yeast YG domain in peroxisomal import (Fig. 1C). After proteolytic removal of the tag (fig. S9) and con- centration to 40 mg/ml (2 mM) in the pres- ence of denaturant, the fragment indeed formed a rigid, translucent gel after several hours (Fig. 6B, left photograph). Higher con- centrations could not be reached because the fragment gelled spontaneously, even when denaturant was included to retard gelation. A more dilute solution at 4 mg/ml (0.2 mM) also formed a hydrogel but after several days. The YG domain is necessary and sufficient for gelation. When all tyrosines in the YG do- main were mutated to serines (Y→S; Fig. 6A), a solution of the corresponding purified frag- ment never formed a gel and remained fluid (Fig. 6B, middle photograph). Furthermore, the isolated YG domain (without the upstream flexible segment; Fig. 6A) gelled as readily as the full-length fragment (Fig. 6B, right photo- graph). Interestingly, the gels liquified after heating to >50°C and resolidified after cooling to room temperature; this cycle could be re- peated many times. The purified protein thus behaves like gelatin, even though the amino acid compositions of the two proteins are very different. This reversible thermosensitivity dif- fers from that of the FG domain of the nu- cleoporin NUP98, whose phase separation is more favorable at higher temperatures (46). The difference might be due to hydrogen bonds between tyrosine hydroxyl groups contribut- ing to the cohesion of the YG hydrogel. PEX5 partitions into YG hydrogels We next examined the permeation properties of the YG hydrogels. Small drops of the N-terminal fragment were gelled at 40 mg/ml in a glass- bottomed dish, and the influx of fluorescently labeled PEX5 or other proteins was imaged on a microscope (Fig. 7A). When buffer contain- ing full-length PEX5 fused to green fluores- cent protein (PEX5-GFP; Fig. 7B) was added, the fusion protein rapidly accumulated at the gel edge and then moved inward (Fig. 7C) at a rate of ~16 nm/s (fig. S10). PEX5-GFP was de- pleted from the buffer just outside the gel, indicating that the rate of entry into the gel is limited only by diffusion. Fluorescence behind the front was essentially constant in the gel, indicating that PEX5-binding sites in this region became saturated. After 30 min of per- meation, the concentration of PEX5-GFP inside the gel was ~20-fold greater than the protein’s concentration in the buffer (Fig. 7D). Although GFP alone could also enter the gel, it showed no enrichment (Fig. 7, C and D). Thus, PEX5 efficiently and specifically partitions into the YG hydrogel, analogously to how NTRs par- tition into hydrogels formed from nucleoporin FG domains. Admission of PEX5 into YG hydrogels re- quires the receptor’s unstructured N terminus. When just the N-terminal region of PEX5 (including all WXXXF/Y motifs) was fused to GFP (Fig. 7B), the resulting fusion protein permeated the gel as rapidly as the full-length version (Fig. 7C) and became similarly enriched (Fig. 7D). By contrast, a fusion between the cargo-binding TPR domain and GFP (Fig. 7B) accumulated to a much lower extent (Fig. 7, C and D). Entry into the gel was driven by the receptor’s WXXXF/Y motifs, because ablating Gao et al., Science 378, eadf3971 (2022) 16 December 2022 6 of 14 RES EARCH | R E S E A R C H A R T I C L E A B unstructured full-length YG AH TM SH3 N + YG N + YG (Y > S) YG alone used for gelation N + YG (Y > S) N + YG YG alone Fig. 6. The YG domain of PEX13 forms hydrogels. (A) Scheme depicts the domain organization of A. thaliana PEX13, which lacks a TM and an SH3 domain (enclosed by dashed lines). The fragments indicated below were expressed in Escherichia coli and purified. Y→S denotes conversion of all tyrosines in the YG domain into serines. (B) Concentrated solutions (40 mg/ml) of the three purified fragments were pipetted into silicone tubing and allowed to gel and then squeezed out of the tubing onto a colored surface and photographed. If gelation had occurred, the solution retained the shape of the tubing; otherwise, the solution remained fluid. Note that the YG domain is necessary and sufficient for gelation. all of the motifs (Fig. 7B) abolished the ac- cumulation of the N-terminal fusion protein inside the gel (Fig. 7, C and D). PEX5 could also drag cargo into the YG hydrogels. GFP fused to a PTS1 import signal (GFP-SKL) showed no enrichment inside the gel on its own but accumulated ~10-fold in the presence of full-length PEX5 (Fig. 7, E to G). The lower partitioning coefficient likely reflects the modest affinity between the import signal and the receptor’s TPR domain (47). The per- meation rate was comparable to that of the linear PEX5-GFP fusion (fig. S10). Partitioning of cargo required the receptor’s WXXXF/Y motifs, because full-length PEX5 that lacked all of the motifs (Fig. 7E) was unable to drag GFP-SKL into the gel (Fig. 7, F and G). Al- though the TPR domain is necessary for the interaction with the import signal, the domain is dispensable for the actual partitioning. Indeed, GFP that lacked the import signal was completely inert in the presence of full-length PEX5 but could be efficiently brought into the gel by a construct consisting of the receptor’s N-terminal region fused to a GFP nanobody (PEX5-GNB; Fig. 7, E and F). In this case, GFP was enriched ~30-fold inside the gel (Fig. 7G), consistent with the higher affinity of the nanobody for the fluorescent protein (48). These results dem- onstrate that the YG domain of PEX13 forms hydrogels which selectively admit PEX5 and PEX5•cargo complexes. Discussion Our results with intact yeast membranes and synthetic hydrogels reveal the existence of a nuclear pore–like conduit on peroxisomes (Fig. 8A). This conduit consists of a dense meshwork—a selective phase—assembled from the YG domains of multiple PEX13 molecules that sit in the membrane in opposite orienta- tions. Similarly to the meshwork formed by cohesive nucleoporin FG domains inside nu- clear pores, the YG phase restricts passage of large soluble molecules. The YG phase can be selectively traversed by the import receptor PEX5, allowing bound cargo to move across the peroxisomal membrane. This mechanism thus allows folded and oligomeric proteins to be imported into peroxisomes, analogously to how folded cargo is moved into and out of the nucleus. The use of tyrosines to form a selective phase is not unprecedented. For example, the protein FUS relies on tyrosine repeats to form hydrogels in vitro and to phase-separate into stress granules in vivo (49). Nucleoporin FG domains with tyrosines in place of phenyl- alanines also retain their ability to form hy- drogels (44). Although these gels no longer accommodate NTRs, they do allow the entry of proteins engineered to bind the modified phase (50). The permeation rate of PEX5 into the per- oxisomal YG hydrogels (~16 nm/s) implies that cargo would traverse the 4-nm peroxisomal membrane in less than a second, which seems physiologically reasonable, although trans- location rates in vivo are not known. PEX5 presumably enters the YG phase by locally dissolving the meshwork, relying predominantly on the WXXXF/Y motifs in its flexible N- terminal region (Fig. 8B). Entry of PEX5 into the YG phase might be regulated by cargo binding, because the association of the recep- tor with peroxisomes has been reported to rely on the presence of cargo (51). The interaction between PEX5 and the YG phase must be weak and transient, allowing the receptor to rapidly bind and dissociate to diffuse through the meshwork. The receptor’s TPR domain also has some affinity for the YG phase and would therefore enhance diffusion, which might be relevant for translocation of larger cargo. However, only the WXXXF/Y motifs are strictly conserved among peroxisomal import recep- tors (1), suggesting that they are sufficient to drive import both of PTS1 and PTS2 proteins. The proposed mechanism implies that peroxisomal import receptors can diffuse through the YG phase in either direction, similarly to how NTRs move bidirectionally through the nuclear pore. Whereas the direc- tionality of nuclear transport is enforced by the Ran•GTP/GDP gradient (GTP, guanosine triphosphate; GDP, guanosine diphosphate), peroxisomal import may be driven instead by the highly favorable interaction between the receptors’ WXXXF/Y motifs and the lumenal domain of the membrane protein PEX14 (52). This interaction might help pull PEX5 out of the YG phase on the lumenal side and would also prevent retrograde diffusion of the receptor back into the cytosol (Fig. 8A). The interaction is likely potentiated by avidity, because import receptors frequently have several WXXXF/Y motifs (53), and PEX14 is known to form oligo- mers that associate with PEX13 (25, 54, 55). Sustained import would thus require unliganded PEX14, which could only be generated by con- tinual retrieval of PEX5 from the lumen by ATP- dependent recycling (Fig. 8A). This model explains why import, per se, is independent of nucleotide hydrolysis and yet occurs against Gao et al., Science 378, eadf3971 (2022) 16 December 2022 7 of 14 RES EARCH | R E S E A R C H A R T I C L E A buffer PEX5 influx gel microscope objective -150 150 0 distance ( m) C PEX5 (full-length) GFP GFP PEX5 (N) GFP B full-length N TPR PEX5 GFP TPR N N + WF TPR GFP GFP WxxxF/Y N AxxxA N PEX5 (TPR) GFP GFP GFP GFP PEX5 (N + WF) GFP buffer side gel side buffer side gel side buffer side gel side buffer side gel side buffer side gel side 1 min 10 min 30 min 40 enrich- ment 1 -150 0 distance ( m) 150 -150 0 distance ( m) 150 -150 0 distance ( m) 150 -150 0 distance ( m) 150 -150 0 distance ( m) 150 D F full-length GFP N TPR N + WF 0.9 4.3 0.9 E 24.4 29.3 PEX5 WF 0 10 20 30 GNB enrichment PEX5 TPR TPR GNB N AxxxA N N N G PEX5 + GFP-SKL PEX5 + GFP-SKL + GFP-SKL + GFP-SKL + GFP-SKL PEX5 ( WF) PEX5 PEX5-GNB 11.6 0.7 0.4 0.0 33.2 0 10 20 30 enrichment PEX5 + GFP-SKL buffer side gel side GFP-SKL buffer side gel side PEX5 ( WF) + GFP-SKL buffer side gel side PEX5 + GFP PEX5-GNB + GFP buffer side gel side buffer side gel side 1 min 10 min 30 min 40 enrich- ment 1 -150 0 distance ( m) 150 -150 0 distance ( m) 150 -150 0 distance ( m) 150 -150 0 distance ( m) 150 -150 0 distance ( m) 150 Fig. 7. The import receptor PEX5 selectively enters YG hydrogels and brings cargo along. (A) YG-hydrogel droplets (40 mg/ml) were prepared in glass-bottomed dishes; permeation of the gels by fluorescently labeled PEX5 or other proteins was imaged by point-scanning confocal microscopy. (B) Scheme depicting the PEX5 fragments that were fused to GFP. PEX5’s N-terminal region contains several WXXXF/Y motifs (magenta arrows), which were mutated to AXXXA in the N + DWF mutant. (C) YG-hydrogel droplets were bathed in buffer containing the indicated GFP-fusion proteins (or GFP alone), and the interface between the buffer and gel was imaged over time. Shown are three selected time points; the fold enrichment of each protein, relative to buffer, across the imaged field is plotted below (mean ± the range of three experiments). (D) Mean enrichment (± the range of three experiments) of the indicated proteins inside the gel compared to buffer after 30 min of imaging as in (C). (E) Nonfluorescent PEX5 variants used for experiments with fluorescent cargo. GNB, anti-GFP nanobody. (F) Same as in (C) except with the indicated PEX5 variants, and GFP with or without the PEX5-binding SKL signal. (G) Mean enrichment (± the range of three experiments) of GFP or GFP-SKL after 30 min of imaging as in (F). Gao et al., Science 378, eadf3971 (2022) 16 December 2022 8 of 14 RES EARCH | R E S E A R C H A R T I C L E A PEX5 cargo PEX13 cytosol PEX14 B Ub PEX1 PEX6 PEX5 lumen PEX2-10-12 import cargo cargo nuclear pore C cytosol nucleus D export NTR import Fig. 8. Model of peroxisomal matrix protein import. (A) PEX13 molecules form a conduit (blue) in the peroxisomal membrane filled with a meshwork (orange) of their YG domains. PEX5 crosses this barrier with bound cargo, using WXXXF/Y motifs (magenta) in its flexible N-terminal region. The interaction between the WXXXF/Y motifs and the lumenal domain of PEX14 retains the receptor inside the organelle. To return to the cytosol, PEX5 is monoubiquitinated (Ub) by the PEX2-10-12 ubiquitin ligase complex and pulled out by the PEX1-PEX6 ATPase. Unfolding of PEX5 during export enables cargo to be released in the lumen. After refolding in the cytosol, and having Ub removed by deubiquitinases, PEX5 can begin another import cycle. (B) PEX5 partitions into the YG meshwork as an extended polypeptide, whose WXXXF/Y motifs locally disrupt the cohesive tyrosine interactions that hold the meshwork together. (C) Diagram illustrating the scaffold (yellow) of the nuclear pore complex, filled with a meshwork of nucleoporin FG domains (orange). Note how the nuclear pore is suspended outside the bilayer instead of being embedded in it like the peroxisomal pore. (D) NTRs use hydrophobic pockets and patches in folded domains to partition into the FG meshwork and diffuse through the nuclear pore. the concentration gradient of the cargo: The energy derives from PEX5 export, because new receptor molecules can only enter the lumen once the previously imported ones have been retrieved. Notably, the proposed model does not contradict a possible role of PEX14 in re- cruiting cargo to peroxisomes (39). The YG meshwork is suspended in the per- oxisomal membrane by PEX13 molecules of opposite orientations. Such dual topology is unusual (56) and may be generated by random insertion of PEX13 directly into the peroxisomal membrane, likely with the involvement of PEX3. The dual topology facilitates assembly of the meshwork by allowing the YG domains to meet and associate within the plane of the membrane. The location of the YG meshwork directly in the lipid bilayer contrasts with the nucleoporin FG meshwork, which forms out- side the nuclear membrane in the aqueous central channel of the nuclear pore complex (compare Fig. 8, A and C). Our hydrogel ex- periments demonstrate that the N-terminal region preceding the YG domain can be ac- commodated in the YG meshwork. This obser- vation explains why the N terminus of PEX13 can cross the membrane and become acces- sible to proteases and modification reagents while the C terminus remains fixed in the membrane in one of the two orientations. Notably, the two orientations reconcile previ- ously conflicting observations regarding the topology of the C terminus (39). The popula- tion of PEX13 with its PEX5-binding SH3 domain exposed to the cytosol could help recruit PEX5 to the organelle. We suspect that the membrane-spanning walls that encircle the YG meshwork are formed by the long AH of PEX13. The AH is the only feature of PEX13, besides the YG domain, that is strictly conserved. Our data show that the AH is also essential for import. Given its length (>60 residues) and dual topology, we surmise that alternating orienta- tions of the AH assemble in a tilted manner into a pore-like structure in the peroxisomal membrane (fig. S11A), with their hydrophilic face oriented toward the aqueous YG phase in the middle and their hydrophobic face toward the lipids on the outside (fig. S11B). The diameter of this structure must be at least 9 nm (the size of the largest cargo reported to enter yeast peroxisomes) (19) but considerably smaller than the nuclear pore (~40 nm in yeast) (57). Indeed, PEX13 has been observed to form nanometer-sized clusters in the per- oxisomal membrane by super-resolution micro- scopy (58), which are potentially consistent with such a conduit. By analogy to the nuclear pore, the concentration of YG domains inside the conduit must be very high (≥200 mg/ml). Our inability to reach such high concentra- tions in vitro likely explains why our YG hydro- gels incompletely exclude inert material. The similarity between peroxisomal import and nuclear transport raises the question of how proteins are targeted to the correct com- partment. A possible answer is that NTRs have evolved to bind the phenylalanine-rich motifs of FG nucleoporins and exhibit low affinity for tyrosine (44). The peroxisomal meshwork might also be denser than the nuclear pore phase, given the shorter distance between aromatic residues in YG domains compared with nucleo- porin FG domains. Whereas peroxisomal import receptors bind the YG phase using unstructured segments (Fig. 8B), NTRs bind nucleoporin FG domains using large folded domains (59) (Fig. 8D). NTRs could thus be sterically impeded from penetrating the denser peroxisomal meshwork and would also have a lower affinity for the YG phase. PEX5 might not be impeded from traversing the nuclear pore, however, because it does leak into nuclei and maintains its steady-state nuclear exclusion only through continuous CRM1-mediated export (60). A similar export pathway does not exist in peroxisomes, which might therefore require a denser permeability barrier to exclude foreign proteins. Nevertheless, the barrier is likely not so tight as to exclude small molecules (35, 40). The limited permeability of the peroxisomal mem- brane might explain why a nuclear pore–like import mechanism, by which a folded protein crosses the membrane, can be accommodated. By contrast, other organelles (such as the ER and mitochondria) need a mechanism that main- tains a tighter seal during protein translocation Gao et al., Science 378, eadf3971 (2022) 16 December 2022 9 of 14 RES EARCH | R E S E A R C H A R T I C L E and thus require imported proteins to be unfolded. Materials and methods All reagents were obtained from Millipore- Sigma unless specified otherwise. Buffers were prepared in ultrapure water. RT denotes room temperature. Plasmid construction The coding sequence of each recombinant protein was codon-optimized for Escherichia coli and inserted into plasmid pET-28b(+) by Gibson Assembly (from New England BioLabs). PEX5 proteins and cargo were produced as fusions containing N-terminal glutathione- S-transferase (GST). A 3C protease–cleavage site was introduced between the GST and the recombinant protein, preceded by the amino acid sequence GSD. The sequence coding for full-length, wild-type PEX5 corresponds to iso- form X3 of the Xenopus laevis PEX5.S gene (GenBank accession no. XP_018082765.1). The fragment corresponding to the N-terminal region spans amino acids 1 to 268, whereas the fragment corresponding to the cargo-binding TPR domain spans amino acids 260 to 579. WXXXF/Y motifs in the N-terminal region were converted to AXXXA by mutagenesis. GFP fusions to PEX5 or to the fragments re- ported in the text were generated by inserting the sequence encoding monomeric enhanced GFP (mEGFP) downstream of each protein, separated by either a GGSGGS linker (for full- length PEX5 or the TPR domain) or a GS linker (for the N-terminal region). GFP-SKL was as- sembled by fusing a modified peroxisomal targeting signal from Photinus pyralis (firefly) luciferase (i.e., the amino acid sequence YKGLKSKL) (47) to the C terminus of mEGFP, preceded by a glycine. GFP alone corresponds to mEGFP without the targeting signal. The plasmid encoding the N-terminal region of PEX5 fused to an anti-GFP nanobody was described previously (10). Fragments encoding the YG domain of PEX13 were produced as fusions to an N-terminal 14× polyhistidine (14×His) tag corresponding to the amino acid sequence SKHHHHSGHHHTGHH- HHSGSHHHTGS. A TEV protease–cleavage site was introduced between the 14×His tag and the recombinant protein. The fragment encompassing the entire unstructured N ter- minus of PEX13 corresponds to amino acids 1 to 193 of PEX13 from A. thaliana (GenBank accession no. NP_187412.1); the YG domain alone corresponds to amino acids 76 to 193. Tyrosines within the YG domain were mu- tated to serines by mutagenesis. All con- structs additionally included a single engineered cysteine at the C terminus for covalent derivatization. Constructs designed for gene deletion and gene expression cassette integration in S. cerevi- siae (yeast) were assembled by restriction-enzyme cloning or Gibson Assembly. The coding se- quence of yeast PEX13 (UniProt accession no. P80667) represents the wild type. Conventional sequences for FLAG and SBP epitope tags, as well as cleavage sites for 3C and TEV proteases, were incorporated into PEX13 at the positions indicated in the text. Constructs containing a C-terminal TEV protease-cleavage site addi- tionally included a diserine linker between the cleavage site and the PEX13 coding se- quence. All cysteine point mutations reported in the text were made by mutagenesis, as well as mutations in the hydrophilic face (mut z: Q150A, E153A, Q164A, and E167A) or the hy- drophobic face (mut ϕ: L155E and L166E) of the AH in PEX13. Conversion of tyrosines in the YG domain of PEX13 (amino acids 71, 81, 86, 90, 95, 101, 108, 110, 114, 115, 119, 123, 127, and 133) into either serines or phenylalanines was performed by de novo gene synthesis (by Integrated DNA Technologies). To replace the YG domain (amino acids 70 to 134) of yeast PEX13, the sequences of YG domains from the following organisms were used: Homo sapiens (human; amino acids 67 to 121, UniProt accession no. Q92968); Dro- sophila melanogaster (fruit fly; amino acids 70 to 133, UniProt accession no. Q7JRD4); A. thaliana (plant; amino acids 79 to 187, UniProt accession no. Q9SRR0); Pichia pastoris (amino acids 25 to 114, UniProt accession no. C4R2I6); Torulaspora delbrueckii (amino acids 59 to 135, UniProt accession no. G8ZRC9); Dictyostelium discoideum (amoeba; amino acids 99 to 214, UniProt accession no. Q54CL3); and Chlamy- domonas reinhardtii (algae; amino acids 70 to 194, UniProt accession no. A0A2K3CZW8). All constructs were verified by sequencing. Yeast strains and culture conditions All yeast strains reported in this study were derived from the S. cerevisiae parental strain UTL7A (MATa, ura3-52, trp1, leu2-3/112). Strains were maintained on yeast extract–peptone- dextrose (YPD) medium (1% w/v yeast extract, 2% w/v peptone, 2% w/v dextrose) with or without the antibiotics hygromycin (250 mg/ml; from Fisher, no. 10687010), geneticin (200 mg/ ml; from Fisher, no. 10131035), or nourseo- thricin (100 mg/ml; from Jena Bioscience, no. AB-101), as appropriate. Genomic deletions and insertions were performed by homologous re- combination using lithium acetate–based trans- formation (61). The UTL7A strain was first derivatized by integrating the expression cassette for the violacein biosynthetic pathway (see below) at the leu2 locus. Briefly, the plasmid pWCD1401 (35) was linearized by the restriction enzyme NotI, and the excised 11.5-kb fragment was purified by agarose gel electrophoresis and used to transform exponentially growing UTL7A cells. Clones were selected on synthetic defined (SD) medium containing 2% w/v glucose and 6.7 g/liter yeast nitrogen base without amino acids (from BD Difco) and supplemented with an amino acid mixture lacking leucine (from Sunrise Science). Correct integration of the cassette was confirmed by polymerase chain reaction (PCR), and the resulting violacein- positive (Vio+) strain was used to generate all subsequent strains and corresponds to the wild type. All gene deletions used the natMX nour- seothricin-resistance cassette as the selective marker. The cassette was amplified from plas- mid pFA6A-natMX (from Addgene, no. 19343) by PCR, using primers that introduced 60–base pair (bp) overhangs corresponding to the 5′ and 3′ untranslated regions immediately upstream and downstream of each gene’s open reading frame, respectively. Clones were selected on YPD medium containing nourseothricin (100 mg/ml), and replacement of each reading frame by the natMX cassette was confirmed by PCR. To generate strains expressing mutant or epitope-tagged versions of the peroxisomal pro- teins described in the text, the natMX marker in the corresponding knockout strain was re- placed with each mutant’s coding sequence, using either the hygMX hygromycin-resistance or the kanMX geneticin-resistance cassette as the selective marker. Briefly, the relevant coding sequence was PCR-amplified either from plasmid pFA6A-hygMX (from Addgene, no. 19342) or pFA6A-kanMX (from Addgene, no. 39296), together with the corresponding antibiotic resistance marker, using primers that introduced 60-bp overhangs as above. Clones were selected on YPD medium containing hy- gromycin (250 mg/ml) or geneticin (200 mg/ml), as appropriate, and correct insertion of the mu- tant reading frame was validated by PCR. Peroxisomal matrix protein import assay A modified violacein biosynthetic pathway (35) was used to quantitatively measure peroxisomal matrix protein import activity in S. cerevisiae cells. Briefly, an expression cassette encoding three enzymes (VioA, VioB, and VioE), which together produce a green pigment, was integrated into the genome along with a leucine auxotro- phic marker as described above. The first two enzymes (VioA and VioB) reside in the cytosol, whereas the third enzyme (VioE) contains a peroxisome targeting signal and is sequestered inside peroxisomes when import is functional. Only when import is compromised does VioE accumulate in the cytosol and the green pig- ment is made. Pigment production is inverse- ly proportional to import efficiency (35). To measure pigment production, strains were cultured overnight in YPD medium at 30°C with shaking. The next morning, cul- tures were diluted 150-fold in 3 ml of freshly prepared synthetic defined (SD) medium lack- ing leucine and cultured at 30°C for a further Gao et al., Science 378, eadf3971 (2022) 16 December 2022 10 of 14 RES EARCH | R E S E A R C H A R T I C L E 48 hours. Cells were collected by centrifugation and resuspended in 300 ml of glacial acetic acid (from Fisher). The cell suspension was transfer- red to 1.5-ml microfuge tubes and incubated for 10 min at 95°C and then mixed by inversion and incubated for a further 10 min. Debris were sedimented by centrifugation at 8000g for 5 min at RT, and the resulting supernatants were transferred to a 96-well, round-bottom black-walled plate (from Corning, no. 3792). Fluorescence of each solution was measured with a Bio-Tek Synergy Neo2 microplate reader, using excitation and emission bands of 535 ± 5 and 585 ± 5 nm, respectively. To calculate rel- ative import activity, the fluorescence reading from each strain was normalized to the average reading from wild-type cells (set to 100%) and pex13D cells (set to 0%) by linear interpolation. Disulfide-mediated cross-linking Yeast cells expressing FLAG-tagged PEX13 with introduced cysteines were cultured overnight in YPD medium at 30°C. Cells were collected by centrifugation, washed once with water, and resuspended in cold lysis buffer (20 mM HEPES•KOH pH 6.8 at RT, 150 mM potassium acetate, 5 mM magnesium acetate, 250 mM sorbitol, and 1 mM EDTA) supplemented with 1 mM phenylmethylsulfonyl fluoride (PMSF) and a protease inhibitor cocktail (from Bimake, no. B14001) according to the manufacturer’s instructions. The cell suspension was transferred to 2-ml screw-capped tubes (from Sarstedt, no. 726994005) on ice and mixed with 0.5-mm prechilled glass beads, and the cells were lysed by bead-beating on a Biospec Products Mini- Beadbeater-16 (no. 607) for four 30-s cycles with 2-min cooling on ice in between. The cell lysate was clarified by low-speed centrifuga- tion at 2000g for 5 min at 4°C to remove intact cells and cellular debris, and the resulting supernatant was recentrifuged at 20,000g for 10 min at 4°C to sediment heavy membranes, including peroxisomes. The final pellet was gently resuspended in assay buffer (20 mM HEPES•NaOH pH 7.2, 50 mM NaCl, 250 mM sucrose, and 1 mM EDTA) by pipeting followed by mixing on a rotator at 4°C for 15 min. The homogeneous membrane suspension was treated with or without the oxidizing agent 4,4′-dithiodipyridine (Aldrithiol-4; from Millipore-Sigma, no. 143057), at the concentra- tions indicated in the text, for 30 min at 30°C with gentle agitation. Where indicated, 0.5% w/v of n-dodecyl-b-D-maltoside (DDM; from Anatrace, no. D310) was added before the oxi- dant to solubilize the membranes. Unreacted cysteines were quenched with 10 mM NEM at 30°C for 5 min. Membranes were then solubi- lized by shaking for 1 hour at 4°C with 0.5% w/v DDM, and the solubilisate was clarified by centrifugation at 20,000g for 10 min at 4°C. PEX13 was immunoprecipitated from the re- sulting supernatant using anti-FLAG M2 aga- rose beads (no. A2220) and eluted by heating in Laemmli buffer as described below. Protease protection Homogeneous membrane suspensions from yeast cells expressing PEX13 with protease- cleavage sites (and FLAG or SBP epitope tags) were prepared as described above. Suspen- sions were mixed with 2 mM homemade 3C protease, 2 mM homemade TEV protease, or both proteases together, as indicated in the text, and incubated for 16 hours at 4°C. Where specified, 10 mM NEM was included in the reactions to preinactivate the proteases. All reactions were ultimately quenched with 10 mM NEM, and the membranes were solubilized with 0.5% w/v DDM for 1 hour at 4°C with shaking. The solubilisate was clarified by centrifugation at 20,000g for 10 min at 4°C, and PEX13 was then immunoprecipitated from the resulting supernatant using either anti-FLAG M2 agarose or streptavidin agarose beads (from Thermo- Fisher, no. 20353), as indicated in the text. Precipitated material was eluted with 0.4 mg/ ml 3×FLAG peptide (from Bimake, no. B23112) or 4 mM biotin, as appropriate, in assay buffer containing 0.05% w/v DDM, before being prepared for SDS-PAGE. For the experiment shown in Fig. 5B, the membrane suspension was first digested with 3C protease as described above and then treated with 200 mM Aldrithiol-4 for 30 min at 30°C to promote disulfide-bond formation. Reactions were quenched with NEM and pro- cessed for SDS-PAGE. PEGmal modification Homogeneous membrane suspensions from yeast cells expressing FLAG-tagged PEX13 with introduced cysteines were prepared as de- scribed above. Suspensions were incubated for 90 min at 4°C and gentle agitation with 2 mM methoxypolyethylene glycol maleimide (PEGmal) of the following sizes (all from Millipore-Sigma), as specified in the text: 10-kD (no. 712469), 5-kD (no. 63187), 2-kD (no. JKA3124), and 0.8-kD (no. 712558). Where indicated, 0.5% w/v DDM was added before PEGmal to solubilize the membranes. Unre- acted PEGmal was quenched with 10 mM cysteine, and PEX13 was immunoprecipitated with anti-FLAG M2 agarose and prepared for SDS-PAGE as described above. Electrophoresis and immunoblotting Unless specified otherwise, samples were heated in Laemmli buffer (from Bio-Rad, no. 1610747) for 5 min at 95°C, with or without 50 mM DTT (from GoldBio), and electrophoretically resolved under denaturing conditions on 4-20% TGX precast polyacrylamide gels (from Bio-Rad). For Coomassie-blue staining, gels were first fixed in 50% v/v methanol and 10% v/v acetic acid for 30 min at RT, then incubated for a further 30 min in 50% v/v methanol, 10% v/v acetic acid, and 0.001% w/v Coomassie Brilliant Blue R-250 (from Bio-Rad), and finally destained overnight in 10% v/v acetic acid before being washed into water. For immunoblotting, proteins were transferred overnight onto nitrocellulose membranes (from Bio-Rad, no. 1620112). Mem- branes were blotted in TBST buffer (20 mM Tris•NaOH pH 7.5, 150 mM NaCl, and 0.1% v/v Tween 20) containing 3% w/v non-fat milk solids (from Apex, no. 20241) using antibodies against the FLAG epitope (no. F7425) or the SBP tag (no. MAB10764), and fluorescently labeled secondary antibodies IRDye 800CW or IRDye 680CW (from LI-COR Biosciences), as appropriate. Blots were imaged on a LI-COR Odyssey M imaging system. Determination of expression levels To validate expression of FLAG-tagged PEX13 constructs reported in this study, membranes from the relevant strains were first solubilized in 0.5% w/v DDM as described above, followed by immunoprecipitation using anti-FLAG M2 beads before processing for SDS-PAGE. Bioinformatic analysis The amino acid composition of the YG domain of PEX13 versus nucleoporin FG domains was calculated by the ProtParam tool on the ExPASy server (62), using the sequences of PEX13, NUP62, and NUP98 homologs from the orga- nisms shown in table S1. The number of amino acids between consecutive aromatic residues (i.e., the spacer length) in YG domains versus nucleoporin FG domains was calculated using a custom script and the sequences of PEX13, NUP62, and NUP98 homologs from the orga- nisms listed in table S1. Because FG repeats consist not only of individual FG motifs but also of FXFG motifs (where “X” denotes any amino acid), such tandem motifs were considered as single aromatic clusters for calculating the spacer length. Protein purification Recombinant proteins were produced in E. coli BL21 Rosetta 2(DE3) cells (from Novagen) by induction with isopropyl b-D-1-thiogalactopyra- noside (IPTG; from GoldBio, no. I2481C50). All PEX5 proteins and cargo were expressed and purified by glutathione-affinity and size-exclusion chromatographies as described previously (10). Briefly, the proteins were eluted from the glutathione resin by proteolytic removal of their GST tag, gel-filtered into 40 mM HEPES•KOH pH 7.8 at RT, 100 mM KCl, 250 mM sucrose, 1 mM MgCl2, and 1 mM DTT, and concentrated to 100 mM before snap-freezing in single-use aliquots. PEX13 fragments containing the YG domain were purified under denaturing conditions. Bacteria transformed with the desired plas- mid were cultured in baffled flasks on an orbital Gao et al., Science 378, eadf3971 (2022) 16 December 2022 11 of 14 RES EARCH | R E S E A R C H A R T I C L E shaker at 37°C in 2×YT medium (from Fisher) containing 50 mg/ml kanamycin and 34 mg/ml chloramphenicol. On reaching an optical den- sity (OD) of 0.6, the cultures were cooled at RT for 30 min and then supplemented with 1 mM IPTG and incubated with shaking at 30°C for an additional 6 hours. Cells were collected by centrifugation at 4000g, rinsed in phosphate- buffered saline, and frozen. Cell pellets were resuspended in freshly prepared lysis buffer (8 M urea, 100 mM sodium phosphate, 10 mM HEPES•NaOH, 10 mM imidazole, 5 mM DTT, and pH adjusted to 8.0 at RT just before use), incubated for 30 min at RT, and then homo- genized by sonication. The cell lysate was clari- fied by centrifugation at 15,000g for 30 min at 20°C, and the resulting supernatant incubated with nickel-charged nitriloacetic acid (Ni-NTA; from ThermoFisher) resin for 1 hour at RT with agitation. Beads were washed with an excess of lysis buffer and then with an excess of gelation buffer (2 M urea, 50 mM HEPES• NaOH, and 1 mM EDTA, pH adjusted to 8.0 at RT just before use) supplemented with 1 mM tris(2-carboxyethyl)phosphine (TCEP; from GoldBio). Bound protein was eluted with 500 mM imidazole in gelation buffer, supple- mented with homemade 6×His-tagged TEV protease, and dialyzed overnight at RT against gelation buffer containing TCEP. The dialyzed solution was clarified by centrifugation and passed over additional nickel resin to remove the released 14×His tag and TEV protease. The final flow-through served as the starting point for gelation. Preparation and photography of YG hydrogels To initiate gelation, each recombinant fragment (dissolved in 2 M urea, 50 mM HEPES•NaOH pH 8.0 at RT, 1 mM TCEP, and 1 mM EDTA), was concentrated to 2 mM on a centrifugal filter device (from Amicon) at 40°C. The resulting solutions were promptly injected into short pieces of Tygon S3 silicone tubing (from Saint- Gobain) that had been plugged at one end with hot-melt adhesive and incubated at RT for several days to complete gelation. The con- tents of each piece of tubing were then squeezed out onto a colored grid and photographed using a Canon EF 75-300 mm f/4-5.6 III zoom lens and a Canon EOS 20D digital single-lens reflex (DSLR) camera configured to maximize spa- tial resolution and minimize digital noise. YG-hydrogel permeation assay A solution of the unstructured N-terminal region (including the YG domain) from A. thaliana PEX13 was concentrated to 2 mM in gelation buffer as described above. Two-microliter drops were promptly spotted on the bottom of multiple wells of a 96-well glass-bottom plate (from Cellvis, no. P96-1.5H-N) and allowed to set for several hours at RT. The resulting gel droplets were equilibrated overnight in excess assay buffer (25 mM HEPES•KOH pH 7.8 at RT, 130 mM KCl), and imaged on a Leica SP8 X point-scanning confocal system, using a DMI6000 inverted microscope and a 20× 0.70 NA HC Plan Apochromat CS air objective. The objective was centered on the boundary between the buffer and gel, and focus was maintained 5 mm above the glass surface by the Leica Adaptive Focus Control (AFC) system. Six initial frames were acquired at a rate of two frames per minute, and then a solution of fluorescently labeled protein (as indicated in the text) in assay buffer was added and the acquisition continued for a further 30 min. All proteins were used at a final concentration of 500 nM. Images were acquired in Leica LAS X software using a pixel size of 1.14 mm2, scan speed of 100 Hz, and pinhole dilated to 1 Airy unit. Fluorescence was excited using a 490-nm bandlet selected from a white- light laser by an acousto-optic tunable filter (AOTF), and a 500- to 700-nm emission band was collected by a hybrid pixel detector (HyD) operating in standard mode without gating and gain set to 100. Imaging parameters were configured to maximize the signal-to-noise ratio while avoiding saturation. All samples intended to be compared were imaged under identical acquisition settings. Image analysis All analysis was performed in ImageJ (63) on the original, unmodified image data using routine functions. For images of immunoblots, band intensities were measured by densitom- etry. Fluorescence images of YG hydrogels were first background-subtracted and corrected for experimental variations in fluorophore con- centration. To estimate the background, six successive images of the buffer-gel interface were acquired before addition of fluorescently labeled protein and then averaged. The result- ing matrix was subtracted, pixel by pixel, from all frames of the time course acquired after ad- dition of fluorescently labeled protein. Background- subtracted time courses were then normalized by mean fluorescence intensity in the buffer. To measure the protein concentration across the buffer-gel interface over time, a rectangu- lar selection 60 mm by 300 mm was centered on the gel edge, and the mean fluorescence in- tensity across this field was measured at all time points using the plot profile function in ImageJ. The intensities were then normalized to the maximum value for each experimental cohort. To calculate the fold enrichment of each protein inside the gel, the mean intensity was measured inside a 15-mm arc that followed the inner contour of the gel edge at the 30-min time point. This value was then divided by the mean intensity in the buffer 150 mm away from the edge. To calculate permeation rates, the total area inside the gel occupied by the relevant permeating species at each time point was identified by intensity-based thresholding. The resulting displacement values were plot- ted as a function of time and fitted to straight lines whose slope corresponded to the perme- ation rate. Data plotting and statistical analysis All experiments were independently per- formed at least three times. Data were plotted in GraphPad Prism (v. 9.3.1), and, where indi- cated, statistical significance was calcula- ted using Student’s two-tailed unpaired t test. Fits to mathematical models were performed by nonlinear least squares regression in Prism. Image processing for publication and figure assembly All images intended to be compared were processed identically. Fluorescence micrographs were first background-subtracted and corrected as described above and then linearly contrast- stretched in ImageJ to the same bit range. Digitized images of immunoblots and stained gels were contrast-stretched to reveal rele- vant bands but avoid clipping of the back- ground. Digital photographs of YG hydrogels were processed using the Camera Raw plugin (v. 14.5.0.1177) in Adobe Photoshop (v. 23.5.1). Helical wheel diagrams were prepared using HeliQuest (64). Secondary structure predic- tions were performed with AlphaFold (65). Figures were assembled for publication in Adobe Illustrator (v. 25.4.1). REFERENCES AND NOTES 1. R. L. M. Jansen, C. Santana-Molina, M. van den Noort, 2. D. P. Devos, I. J. van der Klei, Comparative genomics of peroxisome biogenesis proteins: Making sense of the PEX proteins. Front. Cell Dev. Biol. 9, 654163 (2021). doi: 10.3389/ fcell.2021.654163; pmid: 34095119 I. J. Lodhi, C. F. Semenkovich, Peroxisomes: A nexus for lipid metabolism and cellular signaling. Cell Metab. 19, 380–392 (2014). doi: 10.1016/j.cmet.2014.01.002; pmid: 24508507 3. C. Lismont, I. Revenco, M. Fransen, Peroxisomal hydrogen peroxide metabolism and signaling in health and disease. Int. J. Mol. Sci. 20, 3673 (2019). doi: 10.3390/ijms20153673; pmid: 31357514 4. Y. Fujiki, Peroxisome biogenesis and human peroxisome- deficiency disorders. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 92, 463–477 (2016). doi: 10.2183/pjab.92.463; pmid: 27941306 5. C. Brocard, A. Hartig, Peroxisome targeting signal 1: Is it really a simple tripeptide? Biochim. Biophys. Acta 1763, 1565–1573 (2006). doi: 10.1016/j.bbamcr.2006.08.022; pmid: 17007944 6. G. J. Gatto Jr., B. V. Geisbrecht, S. J. Gould, J. M. Berg, Peroxisomal targeting signal-1 recognition by the TPR domains of human PEX5. Nat. Struct. Biol. 7, 1091–1095 (2000). doi: 10.1038/81930; pmid: 11101887 7. V. C. Kalel, R. Erdmann, Unraveling of the structure and function of peroxisomal protein import machineries. Subcell. Biochem. 89, 299–321 (2018). doi: 10.1007/978-981-13-2233- 4_13; pmid: 30378029 8. D. Pan, T. Nakatsu, H. Kato, Crystal structure of peroxisomal targeting signal-2 bound to its receptor complex Pex7p- Pex21p. Nat. Struct. Mol. Biol. 20, 987–993 (2013). doi: 10.1038/nsmb.2618; pmid: 23812376 9. M. Kunze, The type-2 peroxisomal targeting signal. Biochim. Biophys. Acta Mol. Cell Res. 1867, 118609 (2020). doi: 10.1016/ j.bbamcr.2019.118609; pmid: 31751594 10. M. L. Skowyra, T. A. Rapoport, PEX5 translocation into and out of peroxisomes drives matrix protein import. Mol. Cell 82, 3209–3225.e7 (2022). doi: 10.1016/j.molcel.2022.07.004; pmid: 35931083 Gao et al., Science 378, eadf3971 (2022) 16 December 2022 12 of 14 RES EARCH | R E S E A R C H A R T I C L E 11. V. Dammai, S. Subramani, The human peroxisomal targeting signal receptor, Pex5p, is translocated into the peroxisomal matrix and recycled to the cytosol. Cell 105, 187–196 (2001). doi: 10.1016/S0092-8674(01)00310-5; pmid: 11336669 12. J. W. Zhang, P. B. Lazarow, PEB1 (PAS7) in Saccharomyces cerevisiae encodes a hydrophilic, intra-peroxisomal protein that is a member of the WD repeat family and is essential for the import of thiolase into peroxisomes. J. Cell Biol. 129, 65–80 (1995). doi: 10.1083/jcb.129.1.65; pmid: 7535304 13. P. Feng, M. L. Skowyra, T. A. Rapoport, Structure and function of the peroxisomal ubiquitin ligase complex. Biochem. Soc. Trans. 2022, BST20221393 (2022). doi: 10.1042/ BST20221393; pmid: 36421406 14. P. Feng et al., A peroxisomal ubiquitin ligase complex forms a retrotranslocation channel. Nature 607, 374–380 (2022). doi: 10.1038/s41586-022-04903-x; pmid: 35768507 15. N. B. Blok et al., Unique double-ring structure of the peroxisomal Pex1/Pex6 ATPase complex revealed by cryo- electron microscopy. Proc. Natl. Acad. Sci. U.S.A. 112, E4017–E4025 (2015). doi: 10.1073/pnas.1500257112; pmid: 26170309 16. M. O. Debelyy et al., Ubp15p, a ubiquitin hydrolase associated with the peroxisomal export machinery. J. Biol. Chem. 286, 28223–28234 (2011). doi: 10.1074/jbc.M111.238600; pmid: 21665945 17. C. P. Grou et al., Identification of ubiquitin-specific protease 9X (USP9X) as a deubiquitinase acting on ubiquitin-peroxin 5 (PEX5) thioester conjugate. J. Biol. Chem. 287, 12815–12827 (2012). doi: 10.1074/jbc.M112.340158; pmid: 22371489 18. S. Léon, J. M. Goodman, S. Subramani, Uniqueness of the mechanism of protein import into the peroxisome matrix: Transport of folded, co-factor-bound and oligomeric proteins by shuttling receptors. Biochim. Biophys. Acta 1763, 1552–1564 (2006). doi: 10.1016/j.bbamcr.2006.08.037; pmid: 17011644 19. P. A. Walton, P. E. Hill, S. Subramani, Import of stably folded proteins into peroxisomes. Mol. Biol. Cell 6, 675–683 (1995). doi: 10.1091/mbc.6.6.675; pmid: 7579687 20. N. Wiedemann, N. Pfanner, Mitochondrial machineries for protein import and assembly. Annu. Rev. Biochem. 86, 685–714 (2017). doi: 10.1146/annurev-biochem-060815-014352; pmid: 28301740 21. T. A. Rapoport, L. Li, E. Park, Structural and mechanistic insights into protein translocation. Annu. Rev. Cell Dev. Biol. 33, 369–390 (2017). doi: 10.1146/annurev-cellbio-100616-060439; pmid: 28564553 22. N. Miyata, Y. Fujiki, Shuttling mechanism of peroxisome targeting signal type 1 receptor Pex5: ATP-independent import and ATP-dependent export. Mol. Cell. Biol. 25, 10822–10832 (2005). doi: 10.1128/MCB.25.24.10822-10832.2005; pmid: 16314507 23. P. Ozimek, M. Veenhuis, I. J. van der Klei, Alcohol oxidase: A complex peroxisomal, oligomeric flavoprotein. FEMS Yeast Res. 5, 975–983 (2005). doi: 10.1016/j.femsyr.2005.06.005; pmid: 16169288 24. M. Meinecke et al., The peroxisomal importomer constitutes a large and highly dynamic pore. Nat. Cell Biol. 12, 273–277 (2010). doi: 10.1038/ncb2027; pmid: 20154681 25. P. Lill et al., Towards the molecular architecture of the peroxisomal receptor docking complex. Proc. Natl. Acad. Sci. U.S.A. 117, 33216–33224 (2020). doi: 10.1073/pnas.2009502117; pmid: 33323485 26. M. Monroe-Augustus et al., Matrix proteins are inefficiently imported into Arabidopsis peroxisomes lacking the receptor- docking peroxin PEX14. Plant Mol. Biol. 77, 1–15 (2011). doi: 10.1007/s11103-011-9782-0; pmid: 21553312 27. F. A. Salomons, J. A. Kiel, K. N. Faber, M. Veenhuis, I. J. van der Klei, Overproduction of Pex5p stimulates import of alcohol oxidase and dihydroxyacetone synthase in a Hansenula polymorpha pex14 null mutant. J. Biol. Chem. 275, 12603–12611 (2000). doi: 10.1074/jbc.275.17.12603 pmid: 10777551 28. C. Williams, B. Distel, Pex13p: Docking or cargo handling protein? Biochim. Biophys. Acta 1763, 1585–1591 (2006). doi: 10.1016/j.bbamcr.2006.09.007; pmid: 17056133 29. Y.-T. Kao, K. L. Gonzalez, B. Bartel, Peroxisome function, biogenesis, and dynamics in plants. Plant Physiol. 176, 162–177 (2018). doi: 10.1104/pp.17.01050; pmid: 29021223 30. K. Stein, A. Schell-Steven, R. Erdmann, H. Rottensteiner, Interactions of Pex7p and Pex18p/Pex21p with the peroxisomal docking machinery: Implications for the first steps in PTS2 protein import. Mol. Cell. Biol. 22, 6056–6069 (2002). doi: 10.1128/MCB.22.17.6056- 6069.2002; pmid: 12167700 31. A. Brennand, D. J. Rigden, P. A. M. Michels, Trypanosomes contain two highly different isoforms of peroxin PEX13 involved in glycosome biogenesis. FEBS Lett. 586, 1765–1771 (2012). doi: 10.1016/j.febslet.2012.05.019; pmid: 22641036 32. H. B. Schmidt, D. Görlich, Transport selectivity of nuclear pores, phase separation, and membraneless organelles. Trends Biochem. Sci. 41, 46–61 (2016). doi: 10.1016/j.tibs.2015.11.001; pmid: 26705895 33. K. Ribbeck, D. Görlich, The permeability barrier of nuclear pore complexes appears to operate via hydrophobic exclusion. EMBO J. 21, 2664–2671 (2002). doi: 10.1093/emboj/ 21.11.2664; pmid: 12032079 34. A. Boisson-Dernier, S. Frietsch, T.-H. Kim, M. B. Dizon, J. I. Schroeder, The peroxin loss-of-function mutation abstinence by mutual consent disrupts male-female gametophyte recognition. Curr. Biol. 18, 63–68 (2008). doi: 10.1016/j.cub.2007.11.067; pmid: 18160292 35. W. C. DeLoache, Z. N. Russ, J. E. Dueber, Towards repurposing the yeast peroxisome for compartmentalizing heterologous metabolic pathways. Nat. Commun. 7, 11152 (2016). doi: 10.1038/ncomms11152; pmid: 27025684 36. Y. Elgersma et al., The SH3 domain of the Saccharomyces cerevisiae peroxisomal membrane protein Pex13p functions as a docking site for Pex5p, a mobile receptor for the import PTS1-containing proteins. J. Cell Biol. 135, 97–109 (1996). doi: 10.1083/jcb.135.1.97; pmid: 8858166 37. B. Agne et al., Pex8p: An intraperoxisomal organizer of the peroxisomal import machinery. Mol. Cell 11, 635–646 (2003). doi: 10.1016/S1097-2765(03)00062-5; pmid: 12667447 38. C. Krause, H. Rosewich, A. Woehler, J. Gärtner, Functional analysis of PEX13 mutation in a Zellweger syndrome spectrum patient reveals novel homooligomerization of PEX13 and its role in human peroxisome biogenesis. Hum. Mol. Genet. 22, 3844–3857 (2013). doi: 10.1093/hmg/ddt238; pmid: 23716570 39. A. Barros-Barbosa et al., Membrane topologies of PEX13 and PEX14 provide new insights on the mechanism of protein import into peroxisomes. FEBS J. 286, 205–222 (2019). doi: 10.1111/febs.14697; pmid: 30414318 40. V. D. Antonenkov, R. T. Sormunen, J. K. Hiltunen, The rat liver peroxisomal membrane forms a permeability barrier for cofactors but not for small metabolites in vitro. J. Cell Sci. 117, 5633–5642 (2004). doi: 10.1242/jcs.01485; pmid: 15509867 41. R. Zimmermann et al., Quantitative proteomics and differential protein abundance analysis after the depletion of PEX3 from human cells identifies additional aspects of protein targeting to the ER. Int. J. Mol. Sci. 22, 13028 (2021). doi: 10.3390/ ijms222313028; pmid: 34884833 42. J. P. Wróblewska et al., Saccharomyces cerevisiae cells lacking Pex3 contain membrane vesicles that harbor a subset of peroxisomal membrane proteins. Biochim. Biophys. Acta Mol. Cell Res. 1864, 1656–1667 (2017). doi: 10.1016/ j.bbamcr.2017.05.021; pmid: 28552664 43. R. L. M. Jansen, I. J. van der Klei, The peroxisome biogenesis factors Pex3 and Pex19: Multitasking proteins with disputed functions. FEBS Lett. 593, 457–474 (2019). doi: 10.1002/1873- 3468.13340; pmid: 30776093 44. S. Frey, R. P. Richter, D. Görlich, FG-rich repeats of nuclear pore proteins form a three-dimensional meshwork with hydrogel-like properties. Science 314, 815–817 (2006). doi: 10.1126/science.1132516; pmid: 17082456 45. S. Frey, D. Görlich, A saturated FG-repeat hydrogel can reproduce the permeability properties of nuclear pore complexes. Cell 130, 512–523 (2007). doi: 10.1016/ j.cell.2007.06.024; pmid: 17693259 46. S. C. Ng, D. Görlich, A simple thermodynamic description of phase separation of Nup98 FG domains. bioRxiv 2022.07.26.501556 [Preprint] (2022); doi: 10.1101/ 2022.07.26.501556 47. E. L. Maynard, G. J. Gatto Jr., J. M. Berg, Pex5p binding affinities for canonical and noncanonical PTS1 peptides. Proteins 55, 856–861 (2004). doi: 10.1002/prot.20112; pmid: 15146484 48. A. Kirchhofer et al., Modulation of protein properties in living cells using nanobodies. Nat. Struct. Mol. Biol. 17, 133–138 (2010). doi: 10.1038/nsmb.1727; pmid: 20010839 49. M. Kato et al., Cell-free formation of RNA granules: Low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767 (2012). doi: 10.1016/ j.cell.2012.04.017; pmid: 22579281 50. S. C. Ng et al., Barrier-properties of Nup98 FG phases ruled by FG motif identity and inter-FG spacer length. bioRxiv 2022.10.14.512266 [Preprint] (2022); doi: 10.1101/ 2022.10.14.512266 51. A. M. Gouveia, C. P. Guimarães, M. E. Oliveira, C. Sá-Miranda, J. E. Azevedo, Insertion of Pex5p into the peroxisomal membrane is cargo protein-dependent. J. Biol. Chem. 278, 4389–4392 (2003). doi: 10.1074/jbc.C200650200; pmid: 12502712 52. J. Saidowsky et al., The di-aromatic pentapeptide repeats of the human peroxisome import receptor PEX5 are separate high affinity binding sites for the peroxisomal membrane protein PEX14. J. Biol. Chem. 276, 34524–34529 (2001). doi: 10.1074/jbc.M104647200; pmid: 11438541 53. L. Emmanouilidis, M. Gopalswamy, D. M. Passon, M. Wilmanns, M. Sattler, Structural biology of the import pathways of peroxisomal matrix proteins. Biochim. Biophys. Acta 1863, 804–813 (2016). doi: 10.1016/j.bbamcr.2015.09.034; pmid: 26450166 54. R. Itoh, Y. Fujiki, Functional domains and dynamic assembly of the peroxin Pex14p, the entry site of matrix proteins. J. Biol. Chem. 281, 10196–10205 (2006). doi: 10.1074/jbc. M600158200; pmid: 16459329 55. A. Schell-Steven et al., Identification of a novel, intraperoxisomal pex14-binding site in pex13: Association of pex13 with the docking complex is essential for peroxisomal matrix protein import. Mol. Cell. Biol. 25, 3007–3018 (2005). doi: 10.1128/MCB.25.8.3007- 3018.2005; pmid: 15798189 56. E. Granseth, “Dual-topology: One sequence, two topologies” in Structural Bioinformatics of Membrane Proteins, D. Frishman, Ed. (Springer, 2010), pp. 137–150. doi: 10.1007/978-3-7091- 0045-5_8 57. C. W. Akey et al., Comprehensive structure and functional adaptations of the yeast nuclear pore complex. Cell 185, 361–378.e25 (2022). doi: 10.1016/j.cell.2021.12.015; pmid: 34982960 58. L. P. Crowe, C. L. Wilkinson, K. R. Nicholson, M. T. Morris, Trypanosoma brucei Pex13.2 is an accessory peroxin that functions in the import of peroxisome targeting sequence type 2 proteins and localizes to subdomains of the glycosome. MSphere 5, e00744-19 (2020). doi: 10.1128/mSphere.00744- 19; pmid: 32075879 59. I. V. Aramburu, E. A. Lemke, Floppy but not sloppy: Interaction mechanism of FG-nucleoporins and nuclear transport receptors. Semin. Cell Dev. Biol. 68, 34–41 (2017). doi: 10.1016/j.semcdb.2017.06.026; pmid: 28669824 60. K. Kırlı et al., A deep proteomics perspective on CRM1- mediated nuclear export and nucleocytoplasmic partitioning. eLife 4, e11466 (2015). doi: 10.7554/eLife.11466; pmid: 26673895 61. R. D. Gietz, R. H. Schiestl, High-efficiency yeast transformation using the LiAc/SS carrier DNA/PEG method. Nat. Protoc. 2, 31–34 (2007). doi: 10.1038/nprot.2007.13; pmid: 17401334 62. M. R. Wilkins et al., Protein identification and analysis tools in the ExPASy server. Methods Mol. Biol. 112, 531–552 (1999). pmid: 10027275 63. C. A. Schneider, W. S. Rasband, K. W. Eliceiri, NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9, 671–675 (2012). doi: 10.1038/nmeth.2089; pmid: 22930834 64. R. Gautier, D. Douguet, B. Antonny, G. Drin, HELIQUEST: A web server to screen sequences with specific a-helical properties. Bioinformatics 24, 2101–2102 (2008). doi: 10.1093/ bioinformatics/btn392; pmid: 18662927 65. J. Jumper et al., Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021). doi: 10.1038/ s41586-021-03819-2; pmid: 34265844 66. Y. Gao, M. L. Skowyra, P. Feng, T. A. Rapoport, Protein import into peroxisomes occurs through a nuclear pore–like phase, Mendeley Data (2022); https://doi.org/10.17632/ tskn79gd9p.1. AC KNOWLED GME NTS We thank R. Erdmann (Ruhr-University Bochum) for the S. cerevisiae wild-type strain UTL7A, B. Gardner (University of California, Santa Barbara) for plasmid pWCD1401 encoding the violacein biosynthetic pathway, and J. Jia for the thermosensitivity analysis. We are also grateful to D. Görlich (Max Planck Institute, Göttingen), S. Shao (Harvard Medical School), and members of the Rapoport lab for insightful discussions and helpful comments on the manuscript. Imaging was performed with the help of M. Lowe Ocaña at the Neurobiology Imaging Facility at Harvard Medical School. Funding: This work was funded by National Institutes of Gao et al., Science 378, eadf3971 (2022) 16 December 2022 13 of 14 RES EARCH | R E S E A R C H A R T I C L E Health grant R01GMO52586 (T.A.R.), the Howard Hughes Medical Institute (HHMI) (T.A.R.), HHMI-sponsored Helen Hay Whitney Foundation fellowship award F-1255 (M.L.S.), and HHMI-sponsored Damon Runyon Cancer Research Foundation fellowship award 2354-19 (Y.G.). Author contributions: Y.G. performed the experiments in yeast and with peroxisomal membranes. M.L.S. performed the experiments with hydrogels in vitro. M.L.S. noted the similarity of PEX13’s YG domain with the FG-repeat domain of nucleoporins. P.F. assisted with the experiments. All authors were involved in the experimental design. M.L.S. and T.A.R. wrote a draft of the manuscript with edits from all authors. T.A.R. supervised the project. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data are available in the main text or the supplementary materials; the original unmodified image data are available from Mendeley Data (66). Reagents generated by this study are available from the corresponding author with a completed materials transfer agreement. License information: Copyright © 2022 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse. This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the author-accepted manuscript of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf3971 Figs. S1 to S11 Table S1 MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. Submitted 18 October 2022; accepted 15 November 2022 10.1126/science.adf3971 Gao et al., Science 378, eadf3971 (2022) 16 December 2022 14 of 14
10.1126_science.adf4272
RES EARCH QUANTUM INFORMATION On-demand entanglement of molecules in a reconfigurable optical tweezer array Connor M. Holland1†, Yukai Lu1,2†, Lawrence W. Cheuk1* Entanglement is crucial to many quantum applications, including quantum information processing, quantum simulation, and quantum-enhanced sensing. Because of their rich internal structure and interactions, molecules have been proposed as a promising platform for quantum science. Deterministic entanglement of individually controlled molecules has nevertheless been a long-standing experimental challenge. We demonstrate on- demand entanglement of individually prepared molecules. Using the electric dipolar interaction between pairs of molecules prepared by using a reconfigurable optical tweezer array, we deterministically created Bell pairs of molecules. Our results demonstrate the key building blocks needed for quantum applications and may advance quantum-enhanced fundamental physics tests that use trapped molecules. E ntanglement lies at the heart of quantum mechanics. It is central to the practical advantage provided by quantum devices (1–3) and relevant to understanding the behavior of many-body quantum systems (4). The ability to create entanglement control- lably has been a long-standing experimental challenge. Molecules have been proposed as a promising platform for quantum simulation and quantum information processing because of their rich internal structure and long-lived interacting states (5–8). In the past two dec- ades, much progress has been made in pro- ducing and controlling molecules at ultracold temperatures, both through coherent assem- bly of ultracold alkali atoms (9) and direct laser-cooling (10). Rapid advances have been made in recent years, including the creation of degenerate molecular gases (11, 12), the creation of molecular magneto-optical traps (10, 13–15), high-fidelity detection of single molecules (16–18), and laser-cooling of com- plex polyatomic molecules (19, 20). In addi- tion, coherent dipolar interactions have been observed in bialkali molecules trapped in op- tical lattices (18, 21). A major outstanding challenge to fully real- izing the potential of molecules has been achieving deterministic entanglement with microscopic control. In this work, we real- ized on-demand entanglement between indi- vidual laser-cooled molecules trapped in a reconfigurable optical tweezer array (Fig. 1A). The approach of molecular tweezer arrays (16, 17, 22, 23) combines the microscopic controllability offered by reconfigurable opti- cal tweezer traps (24–28) with the ability to generate entanglement through the electric dipolar interaction between molecules. We specifically made use of effective spin-exchange 1Department of Physics, Princeton University, Princeton, NJ 08544, USA. 2Department of Electrical and Computer Engineering, Princeton University, Princeton, NJ 08544, USA. *Corresponding author. Email: lcheuk@princeton.edu †These authors contributed equally to this work. interactions that arise between rotational states to entangle pairs of molecules into Bell states (Fig. 1B) (29), which are prototypical maximally entangled states of two particles. Our entangle- ment protocol implements an iSWAP gate (8) that, along with site-resolved single-qubit rota- tions achievable in our platform through the optical addressing of individual molecules, ful- fills the requirement for universal quantum computation. Preparing and initializing arrays of laser-cooled molecules Our work starts with single laser-cooled cal- cium monofluoride (CaF) molecules trapped in a dynamically reconfigurable array of opti- cal tweezer traps (17, 23). Through a series of steps involving laser-cooling, optical trapping, and transport, single molecules are transfer- red from a magneto-optical trap into a one- dimensional (1D) array of 37 identical optical tweezer traps with a uniform spacing of 4.20(6) mm (numbers in parentheses are the standard deviation). Because our laser-cooling scheme relies on a closed optical cycle present only for the X2S(v = 0, N = 1) manifold in CaF (30)—where v and N denote the molecular vibrational and rotational state, respectively— the molecules loaded into the tweezers occupy a single rovibrational manifold. To remove the randomness in tweezer oc- cupation, we used a rearrangement approach pioneered in neutral atom experiments (24, 25). We nondestructively detected the tweezer oc- cupations using a variant of L-imaging (31). The empty tweezers were identified then switched off, and the remaining occupied tweezers were then rearranged into the desired 1D pattern. We characterized the rearrangement procedure by measuring the probability of successfully creat- ing uniform arrays and found a single-particle rearrangement fidelity of 97.4(1)%. As shown in Fig. 2A, we were able to create uniform ar- rays up to a size of 16, with a probability >0.6. The rearrangement fidelity was limited by the nondestructive detection fidelity, with mini- mal loss [0.2(10)%] caused by movement of the tweezer traps. ð After rearrangement, we initialized the in- ternal state of the molecules, which were dis- tributed among the 12 hyperfine states in the X2S(v = 0, N = 1) rovibrational manifold. To prepare molecules into a single hyperfine state, we optically pumped molecules into Dj i ¼ X 2S v ¼ 0; N ¼ 1; J ¼ 3=2; F ¼ 2; mF ¼ 2Þ, where J denotes the total angular momentum excluding nuclear spin, F denotes the total angular momentum, and mF denotes its proj- ection onto the quantization axis. Subsequent microwave sweeps along with an optical clean- out pulse transferred the molecules into the target final state ↑j i ¼ X 2S v ¼ 0; N ¼ 1; J ¼ 1=2; F ¼ 0; mF ¼ 0Þ (Fig. 2C). The overall fidelity of preparing molecules in ↑j i was 82.4(11)%. Our preparation sequence ensures that the dominant preparation error is in the form of unoccupied tweezers, with a small contribution coming from molecules prepared ð i ¼ X 2S v ¼ 0; in the incorrect internal state þj N ¼ 0; J ¼ 1=2; F ¼ 1; mF ¼ 1Þ. The state ini- tialization errors come from imperfect micro- wave transfer, polarization impurity of the optical pumping light, and loss caused by heat- ing in the tweezer traps. After state prepara- tion, we measured a molecular temperature of T = 151(10) mK. ð Probing rotational coherence of single molecules To produce entanglement through the dipo- lar interactions between molecules, we re- quired long coherence times compared with the typical interaction timescales of ~10 ms at our tweezer separations. Achieving long coherence times for optically trapped mole- cules has been an ongoing experimental chal- lenge, with steady advances being made. For molecules, different internal states can expe- rience different trapping potentials that, in combination with motion caused by finite temperature, can lead to decoherence. For 1S bialkali molecules, long coherence times of different nuclear spin states have been reported (32). Work using “magic” trapping conditions has also demonstrated extended coherence times between rotational states in both 1S and 2S molecules (18, 33–36). ð Because the effective spin-exchange inter- actions couple different rotational states, we wanted long rotational coherence times be- tween the two interacting states ↑j i and ↓j i ¼ X 2S v ¼ 0; N ¼ 0; J ¼ 1=2; F ¼ 1; mF ¼ 0Þ. Building on previous work in CaF (36), we iden- tified a pseudo-magic trapping condition in which both spin states experience approxi- mately identical trapping potentials. Our pseudo- magic condition takes into account vector and tensor shifts and is achieved by applying a magnetic field orthogonal to the tweezer light polarization at a reduced tweezer depth Holland et al., Science 382, 1143–1147 (2023) 8 December 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E compared with that used for initial loading and imaging. To measure the resulting coherence time, we prepared pairs of tweezer traps in which one trap is empty and the other is occupied by a molecule initialized in ↑j i. We next applied a Ramsey pulse sequence consisting of two p/2 microwave pulses (first pulse along ^x , second A B 1 2 1 2 ... C (1) Load (2) Rearrange (3) Initialize (4) Interact (5) Detect Fig. 1. Laser-cooled molecules in a reconfigurable optical tweezer array. (A) Single CaF molecules trapped in an optical tweezer array are prepared into closely separated tweezer pairs. Molecules in each pair held by separate tweezer traps interact through the long-range electric dipolar interaction electric dipolar interaction leads to dipolar spin-exchange of rotational excitations. (C) Molecules are loaded stochastically, detected nondestructively, and rearranged into the desired 1D configuration. The molecules are then initialized into a single internal state, and the pair separations are reduced to switch on interactions. After specific interaction times, the pairs are separated and detected state-selectively. ^HSE. (B) The pulse along ^n ¼ cos q^x þ sin q^y) separated by a variable free evolution time (Fig. 3A). The remaining fraction of ↑j i molecules, P↑, oscil- lates as a function of q, with the oscillation amplitude directly measuring the coherence. Fitting to an exponential decay curve yields a bare coherence time T (cid:2) 2 of 2.5(3) ms. Add- ing a spin-echo improves the coherence time to T2 = 29(2) ms. Following previous work that explored dipolar interactions of KRb molecules in an optical lattice (21, 37), we implemented the XY8 dynamical decoupling sequence depicted in Fig. 3B and found that the 1/e coherence time was further extended to 215(30) ms (Fig. 3C). This is consistent with our understanding that the bare coherence times are primarily limited by slow (millisec- ond timescale) fluctuations of ambient mag- netic fields. Observing coherent intermolecular interactions Having achieved sufficiently long rotational coherence times, we next set out to observe coherent spin-exchange interactions. The long- range electric dipolar interaction between the molecules gives rise to resonant exchange of rotational excitations between ↑j i and ↓j i. The resulting spin-exchange interaction is described by the Hamiltonian (cid:1) (cid:3) J HSE ¼ (cid:1) 2 ¼ J ^S þ (cid:3) ^S 1 x ^S þ ^S y ^S 2 x þ ^S (cid:3) 1 ^S þ ^S 2 (cid:3) y 1 2 1 2 (cid:3) þ where ^S i , ^S molecule i and i , ^S x i , ^S y i are spin-1/2 operators for (cid:4) (cid:5) D J ¼ 1 r3 1 (cid:3) 3 cos2q′ d2 4pe0 E ↑jd^j↓ with d ¼ being the transition dipole moment, r ¼ r→(cid:6) (cid:6) (cid:6) being the intermolecular sepa- (cid:6) ration, q′ being the angle between r→ and the quantization axis, and e0 being the free space permittivity. Starting with two molecules in a product state, time evolution under ^H SE can lead to entanglement. For example, two mole- cules initially prepared in the product state ↑j i (cid:4) ↓j i become maximally entangled after interacting for a time t = pħ/(2J), where ħ is Planck’s constant h divided by 2p. In our sys- tem, the quantization axis is orthogonal to the intermolecular separation—that is, q′ = 90°. Fig. 2. Tweezer rearrangement and internal state initialization. (A) Probability of creating defect- free molecular arrays through rearrangement. A fit to pn, where n is the array size, gives a single-particle rearrangement fidelity of p = 0.974(1). (B) Example images of defect-free arrays. (C) Optical pumping (orange arrow) prepares molecules in Dj i. Microwave sweeps (dashed green arrows) transfer Dj i molecules to ↑j i. To observe the effect of spin-exchange inter- actions, we first created pairs of ↑j i molecules at an initial separation of 4.20(6) mm, over which interactions were negligible. We next reduced the pair separation to 1.93(3) mm over 3 ms, at which the interaction strength J (J = h × 43 Hz) becomes appreciable on the co- herence timescale. Subsequently, we applied the Ramsey pulse sequence used above with q = 0. To retain long coherence times, the XY8 decoupling pulses were kept on during the free Holland et al., Science 382, 1143–1147 (2023) 8 December 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Single-particle coherence and spin-exchange oscillations. (A) Ramsey pulse sequence used to measure rotational coherence. (Bottom) Bloch sphere diagrams show the action of the various pulses for a molecule initialized in ↓j i. (B) The XY8 dynamical decoupling sequence. (C) Ramsey contrast of noninteracting molecules versus free evolution time t. Green triangles, red squares, and blue circles indicate the cases for which no spin-echo, one spin- echo, and the XY8 sequence is applied, respectively. Exponential fits give coherence times (1/e) of 2.5(3) ms, 29(2) ms, and 215(30) ms, respectively. (Insets) Ramsey fringes with corresponding sinusoidal fits indicated with the dashed lines. (D) Spin-exchange oscillations at a tweezer separation of 1.93(3) mm. Shown are the ↑↑j i populations measured after the Ramsey pulse sequence, P↑↑, as a function of interaction time t, for molecular pairs initialized in i. The solid curve is a fit to a phenomenological model. (Insets) Fluorescence ↑↑j images at the indicated times. (E) Spin-exchange oscillations at separations of 1.26(2) mm (red pentagons), 1.43(2) mm (orange hexagons), 1.60(2) mm (yellow diamonds), 1.68(2) mm (green squares), 1.93(3) mm (blue circles), and 2.35(3) mm (purple triangles). Curves are offset vertically by 0.3 for clarity. (F) The extracted spin-exchange strength J versus pair separation r. The light red band indicates the theoretical prediction taking into account the finite temperature of the molecules and the uncertainty in the electric dipole moment of CaF. The dashed blue curve indicates the prediction without taking into account finite temperature. (Inset) The single-particle loss rate gD versus pair separation r. evolution time. Because the p-pulses in the XY8 sequence leave ^H SE unchanged, spin-exchange interactions are preserved (38). i, the re- sulting state after the Ramsey sequence is given by (cid:10) yj i ¼ ie(cid:3)i Jt For a molecular pair initialized in ↑↑j (cid:7) (cid:8) Jt 4ℏ (cid:7) (cid:8) Jt 4ℏ i þ i cos 4ℏ sin ↓↓j ↑↑j (cid:9) i ð1Þ and P↑↑ oscillates at an angular frequency of J/(2ħ). As shown in Fig. 3D, we observed oscillations of P↑↑, directly revealing the presence of coherent spin-exchange interactions. By varying the pair separation between 1.26(2) and 2.35(3) mm, we verified that the interaction strength J, as ex- tracted from the oscillation frequency, approx- imately scales as 1/r3, as expected for dipolar interactions (Fig. 3F). Experiments with bulk sam- ples of bialkali molecules in optical lattices have observed coherent spin-exchange both through macroscopic measurements (21, 37) and with single-molecule resolution (18), whereas in this work, we observed coherent interactions between individually prepared, laser-cooled molecules. Examining the P↑↑ oscillations in detail, we found that they damp more quickly when the molecules are closer. This could arise from increased molecular loss and reduced single- particle coherence times at close separations. Additionally, the thermal motion of molecules gives rise to disorder in the spin-exchange cou- pling constant J through variations in the in- termolecular separations, leading to damping. At a fixed molecular temperature, this ef- fect increases at closer separations. To deter- mine which damping mechanism is dominant, we first measured single-particle loss rates and found that they increase at close separations (Fig. 3F, inset). We believe that the loss is caused by parametric heating specific to our scheme of generating tweezer traps using an acousto- optical deflector (AOD) and can be circumvented with other tweezer-generation techniques. Mo- lecular loss, however, does not account for all of the observed damping, especially at close sepa- rations. Independently measured single-particle decoherence rates are also insufficient to ex- plain the damping. This leaves finite molecular temperature as the dominant cause of damp- ing at close distances. Simulations that used experimentally measured temperatures revealed damping rates comparable with the observa- tions (supplementary text and fig. S2) (39). Creating and verifying entanglement of molecules Having established the presence of coherent spin-exchange interactions, we next used them to entangle molecules. Specifically, as proposed in (8), time evolution by ^H SE for a specific time of T = pħ/J implements an iSWAP gate, which is maximally entangling. By applying two additional p/2 pulses along the x axis before and after the iSWAP gate, one can convert a molecular pair prepared in ↑↑j i into the Bell Þ, which is max- ð i þ i ↓↓j state yB imally entangled. As a compromise between maximizing J and minimizing heating loss, we chose a pair separation of 1.93(3) mm for creating yB i ¼ 1ffiffi p 2 ↑↑j i. i j j To demonstrate entanglement of molecules, we measured the Bell state creation fidelity F ¼ yB rj jyB i, where r is the experimentally h obtained density matrix. As pointed out in (40), F acts as an entanglement witness, with F > 1=2 indicating two-particle entanglement. To extract F experimentally, we made use of the relation (41) (cid:4) P↑↑ þ P↓↓ þ C (cid:5) F ¼ 1 2 ð2Þ i and ↓↓j where P↑↑ and P↓↓ are the probabilities of mea- i, respectively, and C is the suring ↑↑j i and ↓↓j amplitude of the coherence between ↑↑j i. To measure P↑↑, we separated the two mo- i and subsequently j lecules after preparing yB Holland et al., Science 382, 1143–1147 (2023) 8 December 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Creating and probing Bell pairs. (A) (Top left) The two-tweezer detection probabilities Pij for the Bell pairs. (Top right) Corresponding probabilities when an additional p pulse is applied before measurement. (Bottom) The full probabilities extracted from top left and top right. The SPAM-corrected populations for the ↑j i– ↓j i subsystem are indicated with the green triangles. (B) Probing Bell state coherence through parity oscillations. The coherence C between ↑↑j i and ↓↓j i is obtained from the amplitude of the oscillations in P as a function of q. (Insets) Example images at the maxima and minima of P. (C) SPAM- corrected Bell state fidelity F SPAM versus hold time t at the initial pair creation distance of 1.93(3) mm are indicated with blue circles. Red squares indicate the corresponding data for pairs separated to a larger distance of 4.20(6) mm after creation. (Inset) The contrast CSPAM versus hold time t for the two separations. The extracted 1/e lifetimes are (F SPAM, CSPAM) = [85(5) ms, 61(3) ms] at the Bell creation distance, and [70(16) ms, 57(8) ms] at the larger separation. measured the probability that both tweezers in a pair appear bright. Because our imaging scheme detects only ↑j i molecules, to measure P↓↓ we applied an additional p-pulse before detection (Fig. 4A) to convert molecules from ↓j i to ↑j i. To obtain the coherence envelope C, we measured parity oscillations as follows. We applied a p/2 pulse about a variable axis ^n ¼ cos q^x þ sin q^y after preparing yB i. From the two-tweezer probabilities Pij, we constructed the parity signal P = P11 + P00 – P10 – P01, where 1 and 0 denote a bright or dark tweezer site, respectively. j i and ↓↓j For a general density matrix that includes the possibility of empty tweezers ( ej i) caused by imperfect state preparation, P can display modulation periodic in 2q and q. The 2q mod- ulation is directly related to the coherence be- tween ↑↑j i, whereas the q modulation is related to single-molecule coherences, such as that between e↑j i and e↓j i. The amplitude of the 2q modulation directly gives C. As shown in Fig. 4B, we found that P displays modulation periodic in 2q. The lack of oscillations periodic in q is consistent with the pulse sequence used. Single molecules prepared in e↑j i i or ↓ej effectively experienced a p-pulse, and no single particle coherence was created. From the population and parity oscillation measurements, we obtained a raw Bell state fidelity of F RAW ¼ 0:524ð6Þ. Correcting for detection errors, we obtained a Bell state fidel- ity F ¼ 0:540 7ð Þ. The raw and measurement- corrected fidelities were above 1/2, showing that entanglement was indeed present and created on demand. Correcting additionally for state preparation errors, we obtained a state-preparation and measurement (SPAM)–corrected fidelity of F SPAM ¼ 0:80ð2Þ. In the context of quantum information processing, the SPAM-corrected Bell state fidelity provides an indication of the quality of the iSWAP gate implemented through spin exchange. Nevertheless, full characteriza- tion of our iSWAP gate and a measurement of its fidelity will require full quantum process tomography. With relevance to quantum simu- lation, our measurements of spin-exchange interactions demonstrate the fundamental building block for simulating XY spin models. In particular, the phase of the 2q oscillation in parity measures the relative phase between ↑↑j i in the Bell state, which is sen- sitive to the sign of J. The observed phase of the oscillation shows that J > 0, indicating anti- ferromagnetic spin-exchange interactions. In addition to quantum information processing and quantum simulation, the ability to create entanglement in our system also paves the way toward quantum-enhanced metrology with trapped molecules (42). i and ↓↓j Having demonstrated deterministic creation of Bell pairs, we next probed their lifetime. We measured the Bell state fidelity F SPAM as a function of hold time and found a lifetime of 85(5) ms, which is largely consistent with but somewhat shorter than that expected from uncorrelated single-particle decoherence. We also examined whether the Bell pairs survive when separated to larger distances after their creation. Specifically, we separated the Bell pairs to a distance of 4.20(6) mm over 3 ms after their creation and measured F SPAM as a function of hold time. Within experimental uncertainty, we found a lifetime identical to the case without separation (Fig. 4C). This abil- ity to preserve entanglement while separating molecules could allow one to bypass the lim- ited range of dipolar interactions through movement of molecules and obtain arbitrary connectivity useful for quantum simulation and information processing. Similar abilities to preserve entanglement have recently been demonstrated in atomic tweezer arrays (43). Last, we examined how Bell state fidelities can be improved. The measurement-corrected fidelities were affected substantially by state preparation infidelity, which is caused by a variety of technical imperfections such as in- complete microwave transfers. A detailed ac- counting of errors shows that state preparation fidelities exceeding 0.95 can be achieved (39). Separately, the SPAM-corrected Bell state fi- delity, which reflects the quality of the iSWAP gate, is substantially affected by the finite mo- lecular temperature. To reveal the importance of state preparation fidelity and molecular temperature, we implemented an alternate state preparation procedure (39) that both provides a higher state preparation fidelity [0.85(1)] and a lower molecular temperature [T = 107(5) mK]. The resulting fidelities without state preparation correction were improved to F RAW ¼ 0:608ð14Þ and F ¼ 0:629ð14Þ , re- vealing the importance of state preparation (two-tweezer probabilities and parity oscilla- tion data are provided in fig. S4) (39). In addition, F SPAM improves to 0.863(25) (39), suggest- ing that molecular temperature is a key factor in achieving high-fidelity entanglement. Discussion and Outlook The observed dependence of SPAM-corrected Bell state fidelities on temperature agrees well with numerical simulations, which in- dicate that lowering the molecular temper- atures further by a factor of ~10 could allow fidelities to reach the 0.99 level (39). Such fur- ther cooling could be achieved with methods such as Raman sideband cooling for molecules Holland et al., Science 382, 1143–1147 (2023) 8 December 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E (44, 45). Further improvements are also pos- sible through optimized entanglement schemes. Recent theoretical work that used quantum optimal control has proposed a robust two-qubit entangling gate for CaF molecules in tweezer traps that provides entanglement fidelities ex- ceeding 0.999, under the assumption that they are cooled near their motional ground state (46). We have demonstrated on-demand entan- glement of molecules in a reconfigurable opti- cal tweezer array. Our work was enabled by several advances we have made in controlling laser-cooled molecules, including initialization of defect-free molecular arrays, achievement of long rotational coherence times in tightly focused tweezer traps, and observation of coherent dipolar interactions between indi- vidual laser-cooled molecules. The ability to entangle molecules on demand is a key build- ing block toward simulating quantum spin models, processing quantum information, and performing quantum-enhanced measurements in the emerging platform of molecular tweezer arrays. Specifically for quantum simulation and information processing, the interactions be- tween long-lived molecular states could offer long evolution times and deep circuit depths. With further advances in lowering molecular temperatures, molecular tweezer arrays could potentially provide performance similar to that of established platforms such as Rydberg atom arrays, trapped ions, and superconducting qubits. Our work on entangling molecules on demand, combined with the recent rapid progress in extending laser-cooling to molec- ular species of increasing complexity (19, 20), opens research avenues such as quantum- enhanced precision measurement by using trapped molecules (47) and explorations of molecular collisions (48) and chemical reac- tions with entangled matter. After the initial submission of this work, we became aware of related work reporting dipo- lar spin-exchange and entanglement between molecules in an optical tweezer array (49). RE FERENCES AND NOTES 1. M. A. Nielsen, I. L. Chuang, Quantum Computation and Quantum Information (Cambridge, Univ. Press, 2010). 2. N. Gisin, G. Ribordy, W. Tittel, H. Zbinden, Rev. Mod. Phys. 74, 145–195 (2002). J. Preskill, Quantum 2, 79 (2018). 3. 4. L. Amico, R. Fazio, A. Osterloh, V. Vedral, Rev. Mod. Phys. 80, 517–576 (2008). 5. D. DeMille, Phys. Rev. Lett. 88, 067901 (2002). 6. L. D. Carr, D. DeMille, R. V. Krems, J. Ye, New J. Phys. 11, 7. 055049 (2009). J. A. Blackmore et al., Quantum Sci. Technol. 4, 014010 (2018). 8. K.-K. Ni, T. Rosenband, D. D. Grimes, Chem. Sci. 9, 6830–6838 (2018). 9. K.-K. Ni et al., Science 322, 231–235 (2008). 10. J. F. Barry, D. J. McCarron, E. B. Norrgard, M. H. Steinecker, D. DeMille, Nature 512, 286–289 (2014). 11. L. De Marco et al., Science 363, 853–856 (2019). 12. A. Schindewolf et al., Nature 607, 677–681 (2022). 13. S. Truppe et al., Nat. Phys. 13, 1173–1176 (2017). 14. L. Anderegg et al., Phys. Rev. Lett. 119, 103201 (2017). 15. A. L. Collopy et al., Phys. Rev. Lett. 121, 213201 (2018). 16. L. R. Liu et al., Science 360, 900–903 (2018). 17. L. Anderegg et al., Science 365, 1156–1158 (2019). 18. L. Christakis et al., Nature 614, 64–69 (2023). 19. D. Mitra et al., Science 369, 1366–1369 (2020). 20. N. B. Vilas et al., Nature 606, 70–74 (2022). 21. B. Yan et al., Nature 501, 521–525 (2013). 22. J. T. Zhang et al., Quantum Sci. Technol. 7, 035006 (2022). 23. C. M. Holland, Y. Lu, L. W. Cheuk, Phys. Rev. Lett 131, 053202 (2023). 24. M. Endres et al., Science 354, 1024–1027 (2016). 25. D. Barredo, S. de Léséleuc, V. Lienhard, T. Lahaye, A. Browaeys, Science 354, 1021–1023 (2016). 26. A. Cooper et al., Phys. Rev. X 8, 041055 (2018). 27. M. A. Norcia, A. W. Young, A. M. Kaufman, Phys. Rev. X 8, 041054 (2018). 28. S. Saskin, J. T. Wilson, B. Grinkemeyer, J. D. Thompson, Phys. Rev. Lett. 122, 143002 (2019). 29. J. S. Bell, Physics Physique Fizika 1, 195–200 (1964). 30. B. K. Stuhl, B. C. Sawyer, D. Wang, J. Ye, Phys. Rev. Lett. 101, 243002 (2008). 31. L. W. Cheuk et al., Phys. Rev. Lett. 121, 083201 (2018). 32. J. W. Park, Z. Z. Yan, H. Loh, S. A. Will, M. W. Zwierlein, Science 357, 372–375 (2017). 33. B. Neyenhuis et al., Phys. Rev. Lett. 109, 230403 (2012). 34. F. Seeßelberg et al., Phys. Rev. Lett. 121, 253401 (2018). 35. J. Lin, J. He, X. Ye, D. Wang, Phys. Rev. A 103, 023332 (2021). 36. S. Burchesky et al., Phys. Rev. Lett. 127, 123202 (2021). 37. J.-R. Li et al., Nature 614, 70–74 (2023). 38. J. Choi et al., Phys. Rev. X 10, 031002 (2020). 39. Experimental and simulation methods, additional supporting data, and supplementary text are available as supplementary materials. 40. C. A. Sackett et al., Nature 404, 256–259 (2000). 41. Q. A. Turchette et al., Phys. Rev. Lett. 81, 3631–3634 (1998). 42. L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, P. Treutlein, Rev. Mod. Phys. 90, 035005 (2018). 43. D. Bluvstein et al., Nature 604, 451–456 (2022). 44. L. Caldwell, M. Tarbutt, Phys. Rev. Res. 2, 013251 (2020). 45. Y. Lu, S. J. Li, C. M. Holland, L. W. Cheuk, Raman sideband cooling of molecules in an optical tweezer array. arXiv:2306. 02455v1 [physics.atom-ph] (2023). 46. M. Hughes et al., Phys. Rev. A 101, 062308 (2020). 47. B. L. Augenbraun et al., New J. Phys. 22, 022003 (2020). 48. L. W. Cheuk et al., Phys. Rev. Lett. 125, 043401 (2020). 49. Y. Bao et al., Dipolar spin-exchange and entanglement between molecules in an optical tweezer array. arXiv:2211.09780v1 [physics.atom-ph] (2023). 50. L. W. Cheuk, C. M. Holland, Y. Lu, On-demand entanglement of molecules in a reconfigurable optical tweezer array. Dryad (2023); https://doi.org/10.5061/dryad.j9kd51chh. 51. L. W. Cheuk, C. M. Holland, Y. Lu, On-demand entanglement of molecules in a reconfigurable optical tweezer array. Zenodo (2023); https://doi.org/10.5281/zenodo.8140983. AC KNOWLED GME NTS We thank S. J. Li, W. Bakr’s group, S. Gopalakrishnan, and J. Thompson for fruitful discussions. We also thank W. Bakr, J. Thompson, S. J. Li, and C. Chiu for careful readings of the manuscript. Funding: This work was supported by National Science Foundation grant 2207518, the Alfred P. Sloan Foundation, and Princeton University. Author contributions: Conceptualization: L.W.C. Methodology: C.M.H., Y.L., and L.W.C. Investigation: C.M.H., Y.L., and L.W.C. Visualization: C.M.H., Y.L., and L.W.C. Funding acquisition: L.W.C. Supervision: L.W.C. Writing – original draft: C.M.H., Y.L., and L.W.C. Writing – review and editing: C.M.H., Y.L., and L.W.C. Competing interests: The authors declare that they have no competing interests. Data and materials availability: Data and code are available at Dryad (50) and Zenodo (51). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf4272 Materials and Methods Supplementary Text Figs. S1 to S5 Table S1 References (52–56) Submitted 20 October 2022; resubmitted 30 November 2022 Accepted 28 September 2023 10.1126/science.adf4272 Holland et al., Science 382, 1143–1147 (2023) 8 December 2023 5 of 5
10.1126_science.adf4915
RES EARCH POLLUTION Mortality risk from United States coal electricity generation Lucas Henneman1,2*, Christine Choirat3, Irene Dedoussi4, Francesca Dominici2, Jessica Roberts5, Corwin Zigler2,6 Policy-makers seeking to limit the impact of coal electricity-generating units (EGUs, also known as power plants) on air quality and climate justify regulations by quantifying the health burden attributable to exposure from these sources. We defined “coal PM2.5” as fine particulate matter associated with coal EGU sulfur dioxide emissions and estimated annual exposure to coal PM2.5 from 480 EGUs in the US. We estimated the number of deaths attributable to coal PM2.5 from 1999 to 2020 using individual-level Medicare death records representing 650 million person-years. Exposure to coal PM2.5 was associated with 2.1 times greater mortality risk than exposure to PM2.5 from all sources. A total of 460,000 deaths were attributable to coal PM2.5, representing 25% of all PM2.5-related Medicare deaths before 2009 and 7% after 2012. Here, we quantify and visualize the contribution of individual EGUs to mortality. A ir pollution exposure is associated with adverse health effects and increased risk of death (1–4). Coal electricity-generating units (EGUs), or power plants, are a major contributor to poor air quality (5–7). Coal, historically a relatively inexpensive fuel, is burned to provide electricity worldwide even as the US and other nations continue to de- bate whether it should remain a part of the energy portfolio amid public health and cli- mate concerns. Global coal use for electricity generation is projected to increase (8), and ongoing instability has pushed European na- tions to increase coal use (9, 10). Although coal EGU air pollution emissions have declined in the US in recent decades (11), defining the health burden posed by coal EGUs and the benefits of actions that have reduced EGU emis- sions remains paramount to informing public health, climate, and energy policies in the US (12) and worldwide. Previous studies that quantified the mor- tality burden from coal EGUs in the US (13–18) relied on estimated concentration response functions (CRFs), which assume that fine par- ticulate matter (PM2.5) from coal emissions has the same toxicity as PM2.5 from all sources. How- ever, evidence indicates (19–25) that exposure to sulfur, sulfates, or PM2.5 from coal emissions may be associated with higher relative morbid- ity or mortality risk than that to other PM2.5 con- stituents or PM2.5 from other sources per unit concentration, although uncertainty remains (26, 27). The limited regional (19–22) and tem- poral (23–25) scope of previous studies, along with the lack of availability of coal-specific ex- posure estimates, has hindered the adoption of coal-specific PM2.5 CRFs in mortality burden calculations, likely leading to underestimates of the mortality burden associated with coal EGUs. In addition, previous studies lack tar- geted evidence regarding which coal EGUs are most responsible for increased mortality risk, and this information is needed to inform policies. To estimate the number of deaths associa- ted with exposure to coal PM2.5 from EGUs, we conducted a national-scale study of individual- level health records covering >650 million person-years in the US Medicare population (≥65 years of age) from 1999 to 2016 (unless otherwise noted, populations throughout this study refer specifically to the Medicare popu- lation) (28). We defined “coal PM2.5” as PM2.5 from coal EGU SO2 emissions. We estimated coal PM2.5 using the HYSPLIT with Average Dispersion (HyADS) model, which accounts for date-specific atmospheric transport of PM2.5 to characterize exposure to PM2.5 from individ- ual EGUs (29–32). We used HyADS, a reduced complexity model, to estimate 22 years of ex- posure to coal PM2.5 (from 1999 to 2020) from each of 480 US EGUs. These calculations would have required multiple orders of magnitude more computation time using a typical full- scale chemical transport model. Our study offers the following contribu- tions. First, we estimated and compared mor- tality risk associated with exposure to coal PM2.5 versus total PM2.5 from all sources, showing that previous analyses underestimated the mortality burden from coal EGUs in the US. Second, we calculated the number of deaths linked to each of the 480 coal EGUs, ranking each with respect to its contribution to the mortality burden and tracking its contribu- tion to the overall mortality burden over time amid implementation of emissions controls 1Department of Civil, Environmental, and Infrastructure Engineering, George Mason University Volgenau School of Engineering, Fairfax, VA, USA. 2Department of Biostatistics, Harvard T.H. Chan School of Public Health, Harvard Data Science Initiative, Harvard University, Boston, MA, USA. 3Institute of Global Health, Faculty of Medicine, University of Geneva, Geneva, Switzerland. 4Section Aircraft Noise and Climate Effects, Faculty of Aerospace Engineering, Delft University of Technology, Delft, Netherlands. 5School of Interactive Computing, Georgia Institute of Technology, Atlanta, GA, USA. 6Department of Statistics and Data Sciences, University of Texas, Austin, TX, USA. *Corresponding author. Email: lhennem@gmu.edu Fig. 1. ZIP code–level coal PM2.5 over time. Box plots (median, first, and third quartiles are shown as horizonal lines and outliers as dots) summarize the distribution of ZIP code levels of coal PM2.5. Map areas shown in white do not have ZIP codes. Plots were produced in R using ggplot2; spatial information comes from the USAboundaries package. Henneman et al., Science 382, 941–946 (2023) 24 November 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E and retirements. Third, we documented the spatial distribution of the mortality burden across the US. Results Changes in exposure to coal PM2.5 over time By averaging ZIP (postal) code levels of coal PM2.5 across the conterminous US, we found that the annual average coal PM2.5 declined from 2.34 mg m−3 (range, 0.01 to 8.80) in 1999 to 0.07 mg m−3 (range, 0.00 to 0.39) in 2020 (Fig. 1). Coal PM2.5 was elevated in the eastern US relative to the western US, with annual average concentrations exceeding 4 mg m−3 in multiple ZIP codes in all years from 1999 to 2008. Coal PM2.5 exposure is a combination of emissions from nearby and distant EGUs (figs. S1 and S2). Coal PM2.5 CRF The Medicare dataset contains records of 32.5 million deaths from 1999 to 2016 (table S1), with the annual number of deaths increasing and death rates decreasing across the study period (fig. S3). We found that a 1 mg m−3 increase in annual average coal PM2.5 was associated with a 1.12% increase in all-cause mortality [relative risk (RR): 1.0125; 95% confidence interval (CI): 1.0113 to 1.0137]). This risk is ~2.1 times greater than the RR associated with exposure to PM2.5 from any source (1.0060 per mg m−3; 95% CI: 1.0053 to 1.0067), which was estimated by Wu et al. in 2020 in the same Medicare cohort using an analogous statistical model (4). Number of excess deaths attributable to coal PM2.5 For each year from 1999 to 2020, we estimated the excess number of deaths attributable to coal PM2.5 relative to what would have occurred assuming zero SO2 emissions from coal EGUs (i.e., coal PM2.5 = 0). Summing over the study period, we estimated that 460,000 (95% CI: 420,000 to 500,000) deaths would have been avoided if all coal EGU SO2 emissions were eliminated (Fig. 2 and table S2). Annual ex- cess deaths attributable to coal PM2.5 were highest between 1999 and 2007, averaging more than 43,000 deaths per year for a total of 390,000 (95% CI: 360,000 to 430,000). After 2007, annual excess deaths declined substan- tially, reaching 1600 (95% CI: 1400 to 1700) in 2020. The total number of deaths in the Medi- care population for the period 1999 to 2020 was 38.6 million (we projected annual deaths in each ZIP code for the period 2017 to 2020 as the average from 2014 to 2016; fig. S3). There- fore, Medicare deaths associated with coal PM2.5 exposure represent 1.2% (95% CI: 1.1 to 1.3%) of all Medicare deaths. Changes in base- line mortality rates had a much smaller influence than changes in coal PM2.5 on the variability in annual deaths associated with coal PM2.5 since 1999 (figs. S4 and S5). Fig. 2. Annual number of excess deaths attributable to coal PM2.5, estimated using the RR for coal PM2.5 from this study and RRs for total PM2.5 from the literature. All excess deaths are estimated relative to zero coal PM2.5. The area filled by horizontal hashing indicates deaths estimated using RRs derived from this study (bounds represent 95% CI). Areas filled by vertical and diagonal hashing correspond to deaths estimated using RRs for total ambient PM2.5 exposure from the literature (4, 33). The gray shaded region from 2017 to 2020 represents years for which ZIP code–specific baseline death rates were assumed from the 2014 to 2016 average. This figure was produced in R using ggplot2. The estimated RR for coal PM2.5 from the statistical model was higher than the previous- ly estimated RRs for total PM2.5 exposure that are often used for risk assessments, implying that the number of excess deaths attributa- ble to coal EGUs was underestimated in prior studies (13–18). For example, by combining coal PM2.5 exposure with two RRs for total PM2.5 previously used in risk assessments, 1.0060 per 1 mg m−3 (95% CI: 1.0053 to 1.0067) estimated for the Medicare population (4) and 1.6% per 10 mg m−3 (95% CI: 1.4 to 1.8) esti- mated for the general population (33), we esti- mated 240,000 (95% CI: 220,000 to 260,000) and 200,000 (95% CI: 130,000 to 280,000) excess deaths from coal EGUs, respectively (Fig. 2). We compared mortality from coal PM2.5 es- timated in the main analysis with an aggre- gate health burden associated with total PM2.5 from all sources. Using the RR reported by Wu et al. (4) for the Medicare population and the same annual PM2.5 exposure used in that analy- sis, we calculated 2,000,000 excess deaths due to ambient PM2.5 from 2000 to 2016 relative to a PM2.5 concentration of 0 (a portion of these excess deaths is attributable to natural emis- sions sources). Thus, our estimates imply that exposure to coal PM2.5 was associated with 25% of all PM2.5-related Medicare deaths from 2000 to 2008 and with 7% of all PM2.5 deaths from 2013 to 2016 (fig. S6). Individual EGU contributions to mortality burden We identified 138 of the 480 coal EGUs that were associated with >1000 excess deaths across the study period and 10 EGUs that were asso- ciated with >5000 deaths (Fig. 3). Although EGUs east of the Mississippi River were asso- ciated with the greatest numbers of deaths because of their high emissions and proxim- ity to population centers, each geographical region contained at least one EGU associated with >400 deaths. The distribution of EGU- specific deaths was heavily skewed; 91% of the total deaths were associated with EGUs that accounted for 50% of nationwide coal EGU SO2 emissions during the study period. Nor- malizing excess deaths by energy produced may rank EGUs differently. Figure 4 shows the temporal trend in the number of deaths associated with each EGU, highlighting the two most harmful coal EGUs within each region. Large declines in the num- ber of deaths corresponded with SO2 emission control installations and facility retirements. For example, for the Keystone facility in Pennsylvania, the average annual number of attributable deaths was 640 (95% CI: 580 to 700) before 2008, but declined to 80 (95% CI: 70 to 90) after scrubber installations in 2009 to 2010. We developed an interactive tool to examine individual EGUs and their contri- butions to state-specific Medicare deaths in relation to SO2 emissions control installations and unit retirements (34). Sensitivity of results to unmeasured confounding The stratified Poisson regression for estimat- ing the CRF was chosen based on its use in previous health impact studies of exposure to total PM2.5 in the Medicare population. The log-linear CRF implied by the model was chosen to facilitate the source-specific attribution of health impacts, but it may not reflect the true Henneman et al., Science 382, 941–946 (2023) 24 November 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Excess deaths associated with individual coal EGUs from 1999 to 2020. EGUs (N = 480) are organized by region to improve interpretability, and the facilities associated with the most deaths are labeled. Inset: total SO2 emissions by location from 1999 to 2020 (hexagonal grids may include multiple EGUs) and regional boundaries. Plots were produced in R using ggplot2; spatial information comes from the USAboundaries package. relationship between coal PM2.5 and mortality risk across all exposure levels during the study period. Although the stratified Poisson model adjusts for many confounders and has been shown in the context of total PM2.5 exposure to be robust to a variety of strategies for con- founding adjustment, we cannot rule out the possibility that unmeasured factors related to mortality risk vary systematically with coal PM2.5 in a manner not captured by observed characteristics in the model. Using the E-value (34, 35), we found that a potential confounder would need to have an association with both mortality rate and coal PM2.5 of 1.125 (lower confidence interval: 1.118) on the RR scale to explain away the association between mortal- ity and coal PM2.5. To explore the potential confounding by air pollution sources other than coal PM2.5, we performed several additional sensitivity analy- ses. We present coal PM2.5 RRs from models that adjust for total PM2.5, residual PM2.5 (total PM2.5 minus coal PM2.5) as a marker for all other sources, NO2 as a marker for primary traffic-related air pollution, and both NO2 and residual PM2.5 (table S3). Adjusting for total PM2.5 attenuated the risk of coal PM2.5 sub- stantially, which is consistent with coal PM2.5 being captured by the total PM2.5 metric. When including residual PM2.5 and/or NO2, we found a slight attenuation in RR from the main mod- el, and the RR for coal PM2.5 remained higher than the RR for total PM2.5 found by Wu et al. (4). Including markers for other PM2.5 sources as confounders introduced important limita- tions, as explained in the supplementary text. Furthermore, we implemented a “first-differences” analysis of within–ZIP code changes over time to adjust for observed and unobserved differ- ences across ZIP codes (fig. S7). This analysis addresses possible threats to validity caused by confounding differences across different locations, providing strong evidence that areas experiencing larger decreases in coal PM2.5 also experienced larger decreases in mortality rates. Results from this analysis support the validity of the primary analysis to quantify the mortality burden with a relative risk adjusted for individual- and ZIP code–level confounders measured throughout the entire study period. Sensitivity of results to HyADS characterization of coal PM2.5 HyADS rescales air parcel location counts ex- tracted from HYSPLIT to coal PM2.5 using a single year’s chemical transport model output, which may introduce errors. Our comparisons (30) of coal PM2.5 with coal PM2.5 source im- pacts from year 2006 Hybrid CMAQ-DDM, a full form model (FFM) bias corrected against observations, confirmed that the spatial pat- tern is well captured and that error and bias are within the typical range of FFMs (although the bias-corrected CMAQ-DDM itself has un- certainty). Because we expected potential errors in coal PM2.5 to be smaller in years surrounding the year when the scaling was performed, we retrained the Poisson regression model three times using data only from subsets of the total years available (1999 to 2003, 2004 to 2007, and 2008 to 2016). The estimated RRs from coal PM2.5 were comparable but slightly larger than in the main analysis from the 1999 to 2003 and 2004 to 2007 models, with a more pro- nounced difference in RR from the 2008 to 2016 model (table S3). The change in RR across Henneman et al., Science 382, 941–946 (2023) 24 November 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Total annual excess deaths associated with each of the coal EGUs in each region, with the two most harmful facilities highlighted. Scrubber installations designate the earliest year that a scrubber was installed at one or more of each EGU’s units [facility information from (46)]. Plots were produced in R using ggplot2. different periods may be consistent with either genuine changes in risk or deterioration of HyADS’ performance in years further from the year on which the scaling was based (2005). To explore sensitivity to the process for scal- ing HyADS to coal PM2.5, we estimated the RR and corresponding excess deaths from coal EGUs using unscaled air parcel counts for each ZIP code output by HyADS and found similar estimates of attributable deaths (650,000; 95% CI: 590,000 to 710,000). Comparing coal PM2.5 estimates against observed sulfate PM2.5 at rural monitors indicated that HyADS may have underestimated exposure and exaggerated ex- posure declines during the study period (a por- tion of this decline in coal PM2.5 is attributable to decreasing EGU contributions to total US SO2 emissions). A sensitivity analysis using a sulfate-adjusted coal PM2.5 metric (fig. S8) es- timates a mortality RR of 1.0147 (95% CI: 1.0135– 1.0158) and 790,000 (95% CI: 720,000–850,000) excess deaths. These findings indicate that, to the extent that HyADS might underestimate coal-derived PM2.5 and exaggerate exposure declines, it provides a conservative estimate of the mortality burden associated with exposure to SO2 emissions from coal EGUs. Future studies may use newly developed approaches for esti- mating CRFs that account for uncertainty in air pollution exposure (36). We used SO2 emissions from coal to derive coal PM2.5 because of evidence that secondary PM2.5 from SO2 emissions constitutes most of the ambient PM2.5 from coal EGUs during the study period (13, 17, 37). Because SO2 emissions and related atmospheric physical-chemical processes that increase ambient PM2.5 are cor- related with complementary processes of other species, e.g., primary PM2.5 and NOx, coal PM2.5 captures the influence of these other species. Although primary PM2.5 emissions are not mea- sured at each EGU, estimated nationwide an- nual primary PM2.5 EGU emissions are correlated (R2 = 0.97) with measured nationwide annual SO2 emissions (38). Sensitivity analyses using observed sulfate and comparisons with alter- native modeling strategies revealed broad con- sistency with the primary analysis, particularly in EGU relative rankings by excess deaths, in- dicating bounds on uncertainties associated with the diversity of technologies and assump- tions available for assessing exposure to EGU SO2 emissions. Comparison with deaths estimated using a chemistry-transport air quality model Although it is impossible to directly validate the estimated number of excess deaths attrib- utable to coal PM2.5, we compared our results with analogous coal EGU health burdens de- rived using atmospheric sensitivities from an FFM. Using GEOS-Chem adjoint PM2.5 sensi- tivities (13) and the coal PM2.5 RR from the main analysis, we estimated 20,000 (95% CI: 19,000 to 22,000) and 13,000 (95% CI: 12,000 to 14,000) excess deaths in 2006 and 2011, re- spectively (these years were chosen to span emissions reductions after 2006 and to align with previously published GEOS-Chem adjoint results). These values are comparable, although smaller (especially in 2006) than the excess deaths estimated from coal PM2.5 exposure in this study of 35,000 (95% CI: 32,000 to 38,000) and 15,000 (95% CI: 13,000 to 16,000) in 2006 and 2011, respectively. Correlations between the number of deaths assigned to each coal EGU by HyADS and GEOS-Chem adjoint were high (R2 ≥ 0.85) for all EGUs and for EGUs in most regions (fig. S9 and table S4), and the two models rank ordered EGUs similarly by their associated deaths (fig. S10). Mean differences in nationwide HyADS EGU-specific death esti- mates relative to the chemical transport model were higher in 2006 (71% for all EGUs) than in 2011 (15%). Although GEOS-Chem adjoint re- sults from the 2 years available are difficult to project to all 22 years of this study, and be- cause chemical transport models, including GEOS-Chem, have uncertainties due to potential bias in emissions inputs, model parameter- izations, and meteorology, agreement between the models at levels consistent with previous studies (39) increases confidence in the results reported here. Implications We conducted the longest-term national study to date estimating the excess number of deaths associated with exposure to SO2 emissions from US coal EGUs. A key innovation in this study is the combined use of coal EGU-specific expo- sure estimates and individual-level health data on the same population during the same time period to estimate the mortality burden. This approach has been hampered until now by the limited availability of large-scale health data- bases and source-specific exposure estimates. Our approach illustrates the utility of deriving air pollution exposure with a combination of dispersion-based and chemical transport models in epidemiological and risk assessment for well- characterized sources. We found that, over the past two decades in the US, coal PM2.5 was associated with 460,000 extra deaths, constituting >22% of total excess deaths attributable to PM2.5. We also found that the mortality burden of coal PM2.5 has been underestimated using traditional impact assessments that rely on CRFs for total PM2.5 mass (13, 16–18, 39–41). The elevated mortality RR associated with annual exposure to coal PM2.5 aligns with previous evidence of in- creased relative health risks associated with coal- related PM2.5 or sulfur or sulfate exposure per unit concentration (19–25), although other studies have found little evidence of increased risk Henneman et al., Science 382, 941–946 (2023) 24 November 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E related to secondary sulfate PM2.5 or PM2.5 associated with coal (26, 27). Large decreases in annual deaths across the study period high- light the success of emissions reductions brought about by regulations under the 1990 Clean Air Act Amendments. Although coal use in the US has remained low, global use is expected to increase and plateau by 2025 (8), suggesting the potential for high mortality costs from coal for years to come. We used SO2 emissions from coal to derive coal PM2.5; however, we cannot conclude that the portion of ambient PM2.5 associated with SO2 emissions emitted from coal power plants is more or less harmful than ambient PM2.5 from other species emitted from coal power plants. Disentangling the mortality risks of the various PM2.5 species emitted from coal EGU emissions is not possible within our mod- eling framework because of the high correla- tion between species emitted from coal EGUs such as NOx and primary PM2.5. Given how we estimate exposure to “coal PM2.5,” our find- ing of a higher mortality risk of exposure to coal PM2.5 relative to other PM2.5 suggests the potential for population health benefits of re- ducing SO2 emissions from coal power plants, for example, by installing emissions control devices or shutting coal facilities completely. Full separation of the health impacts of var- ious emitted species from coal EGUs is of ad- ditional interest to policy-makers because of the varying technologies available to reduce EGU emissions of specific pollutants, and it should be considered in future studies. HyADS benefits from well-characterized source locations and emissions, along with the rela- tively slow atmospheric transformation of emitted SO2 to particulate sulfate. Expanded incorporation of information from observa- tion and chemical transport model–based source apportionment techniques in reduced com- plexity models may enable linkages between emitted species beyond SO2, atmospheric pro- cesses, exposure, and health outcomes. Although source-specific PM2.5 cannot be directly mea- sured, observation-based receptor methods for PM2.5 source apportionment (42) could pro- vide an approximate ground truth (albeit with their own uncertainties) for evaluating mod- eled source-specific exposure. Advanced sensi- tivity approaches incorporated within chemical transport models, such as GEOS-Chem Adjoint used here, and sensitivity methods such as the direct decoupled method (DDM) (43) or the integrated source apportionment method (ISAM) (44) offer model-based approaches that more explicitly incorporate atmospheric chemistry and physics. Expanding computational capac- ity will make comparisons with these types of models in applications with many sources in- creasingly feasible. These results advance the growing body of evidence showing varying toxicity of PM2.5 orig- inating from different sources. Although the US and other countries continue to regulate total ambient PM2.5 concentrations, entities such as the EPA Clean Air Scientific Advisory Commit- tee have specifically cited a need for research to assess health effects associated with changes in PM2.5 composition and sources over time as an important consideration for future PM2.5 policy assessments (45). Our findings have implica- tions for current air pollution risk assessments, which incorrectly assume equal toxicity for am- bient PM2.5 from all sources and for all loca- tions. The research platform that we used to quantify exposure associated with individual coal EGUs, which accounts for pollution trans- port and location relative to population centers, can support more efficient regulatory efforts by producing targeted evidence of how indi- vidual EGU sources contribute to the existing health burden. RE FERENCES AND NOTES 1. D. W. Dockery et al., N. Engl. J. Med. 329, 1753–1759 (1993). 2. F. Laden, J. Schwartz, F. E. Speizer, D. W. Dockery, Am. J. Respir. Crit. Care Med. 173, 667–672 (2006). 3. Q. Di et al., N. Engl. J. Med. 376, 2513–2522 (2017). 4. X. Wu, D. Braun, J. Schwartz, M. A. Kioumourtzoglou, 5. 6. F. Dominici, Sci. Adv. 6, eaba5692 (2020). J. M. Godowitch, G. Pouliot, S. Trivikrama Rao, Atmos. Environ. 44, 2894–2901 (2010). J. A. de Gouw, D. D. Parrish, G. J. Frost, M. Trainer, Earths Futur. 2, 75–82 (2014). 7. L. Xu et al., Proc. Natl. Acad. Sci. U.S.A. 112, 37–42 (2015). 8. International Energy Agency, “Coal 2022: Analysis and forecast to 2025” (IEA, 2022); https://www.iea.org/reports/coal-2022. 9. S. Meredith, “Russia is squeezing Europe’s gas supplies, sparking a bitter and reluctant return to coal,” CNBC, 21 June 2022; https://www.cnbc.com/2022/06/21/ukraine-war-europe-turns- to-coal-as-russia-squeezes-gas-supplies.html. 10. M. Eddy, “Germany will fire up coal plants again in an effort to save natural gas,” The New York Times, 19 June 2022; https:// www.nytimes.com/2022/06/19/world/europe/germany- russia-gas.html. I. C. Dedoussi et al., Environ. Res. Lett. 14, 094003 (2019). 11. 12. US Environmental Protection Agency, “New source performance standards for greenhouse gas emissions from new, modified, and reconstructed fossil fuel-fired electric generating units; emission guidelines for greenhouse gas emissions from existing fossil fuel-fired electric generating units; and repeal of the Affordable Clean Energy Rule” (EPA, 2023); https://www.epa.gov/ system/files/documents/2023-05/FRL-8536-02-OAR% 20111EGU%20NPRM%2020230504_Admin.pdf. I. C. Dedoussi, S. D. Eastham, E. Monier, S. R. H. Barrett, Nature 578, 261–265 (2020). 13. 14. M. P. S. Thind, C. W. Tessum, I. L. Azevedo, J. D. Marshall, Environ. Sci. Technol. 53, 14010–14019 (2019). 15. J. J. Buonocore, X. Dong, J. D. Spengler, J. S. Fu, J. I. Levy, Environ. Int. 68, 200–208 (2014). 16. F. Caiazzo, A. Ashok, I. A. Waitz, S. H. L. Yim, S. R. H. Barrett, Atmos. Environ. 79, 198–208 (2013). I. C. Dedoussi, S. R. H. Barrett, Atmos. Environ. 99, 610–617 (2014). 17. 18. J. Lelieveld et al., Proc. Natl. Acad. Sci. U.S.A. 116, 7192–7197 (2019). 19. P. K. Hopke et al., Environ. Res. 181, 108912 (2020). 20. T. To et al., Environ. Res. 199, 111302 (2021). 21. B. Ostro et al., Environ. Health Perspect. 123, 549–556 (2015). 22. Y. Wang et al., Environ. Int. 158, 106969 (2022). 23. L. R. F. Henneman, C. Choirat, A. C. M. Zigler, Epidemiology 30, 477–485 (2019). 24. S. Weichenthal et al., Environ. Health Perspect. 129, 107005 (2021). 25. G. D. Thurston et al., Environ. Health Perspect. 124, 785–794 (2016). 26. Z. A. Pond et al., Environ. Sci. Technol. 56, 7214–7223 (2022). 27. D. Q. Rich et al., Environ. Int. 126, 387–394 (2019). 28. Centers for Medicare & Medicaid; https://www.cms.gov/. 29. J. A. Casey et al., Nat. Energy 5, 398–408 (2020). 30. L. R. F. Henneman et al., J. Expo. Sci. Environ. Epidemiol. 31, 654–663 (2021). 31. L. R. F. Henneman, C. Choirat, C. Ivey, K. Cummiskey, C. M. Zigler, Atmos. Environ. 203, 271–280 (2019). 32. L. R. Henneman, L. J. Mickley, C. M. Zigler, Environ. Res. Lett. 14, 115003 (2019). 33. D. Krewski et al., Res. Rep. Health Eff. Inst. 140, 5–114, discussion 115–136 (2009). 34. T. J. VanderWeele, P. Ding, Ann. Intern. Med. 167, 268–274 (2017). 35. M. B. Mathur, P. Ding, C. A. Riddell, T. J. VanderWeele, Epidemiology 29, e45–e47 (2018). 36. K. P. Josey, P. deSouza, X. Wu, D. Braun, R. Nethery, J. Agric. Biol. Environ. Stat. 28, 20–41 (2023). 37. C. W. Tessum et al., Proc. Natl. Acad. Sci. U.S.A. 116, 6001–6006 (2019). 38. US Environmental Protection Agency, “Air pollutant emissions trends data: State average annual emissions trend” (EPA, 2018); https://www.epa.gov/air-emissions-inventories/ air-pollutant-emissions-trends-data. 39. E. A. Gilmore et al., Environ. Res. Lett. 14, 074016 (2019). 40. S. L. Penn et al., Environ. Health Perspect. 125, 324–332 (2017). 41. N. Fann, C. M. Fulcher, K. Baker, Environ. Sci. Technol. 47, 3580–3589 (2013). 42. P. K. Hopke, J. Air Waste Manag. Assoc. 66, 237–259 (2016). 43. S. L. Napelenok, D. S. Cohan, Y. Hu, A. G. Russell, Atmos. Environ. 40, 6112–6121 (2006). 44. R. H. F. Kwok, K. R. Baker, S. L. Napelenok, G. S. Tonnesen, Geosci. Model Dev. 8, 99–114 (2015). 45. Clean Air Science Advisory Committee, “CASAC review of the EPA’s policy assessment for the reconsideration of the National Ambient Air Quality Standards for Particulate Matter” (EPA, 2022), report EPA-CASAC-22-002; https://casac.epa. gov/ords/sab/f?p=113:12:1342972375271:::12. 46. US Environmental Protection Agency, “Clean air markets program data” (EPA, 2016); https://campd.epa.gov/. 47. Online data exploration tool “Coal Pollution Impacts Explorer” for: S. Jin et al., Mortality risk from United States coal electricity generation, GitHub (2023); https://cpieatgt. github.io/cpie/. 48. Code for: S. Jin et al., Mortality risk from United States coal electricity generation, GitHub (2023); https://github.com/ NSAPH-Projects/coal_pm25-mortality-medicare. 49. Coal PM2.5 source impacts and coal EGU information for: S. Jin et al., Mortality risk from United States coal electricity generation, (OSF, 2021); https://doi.org/10.17605/ OSF.IO/8GDAU. 50. B. Sabath, “National Causal Analysis Health Outcomes: SAS code to create analytic dataset, GitHub (2021); https:// github.com/NSAPH/National-Causal-Analysis/tree/master/ HealthOutcomes. AC KNOWLED GME NTS We thank S. Jin of the Georgia Institute of Technology for work on the interactive data visualization development. Funding: This work was supported by the National Institutes of Health (grant NIHR01ES026217 to C.Z. and grants R01MD012769, R01ES028033, 1R01ES030616, 1R01AG066793, 1R01MD016054-01A1, 1R01ES 034373-01, 1RF1AG080948, and 1R01ES029950 to F.D.); the US Environmental Protection Agency (EPA) (grant 835872 to C.Z., F.D., and L.H.); EmPOWER Air Data Challenge (L.H., C.Z., and J.R.); the Alfred P. Sloan Foundation (grant G-2020-13946 to F.D.); and The Health Effects Institute (HEI) (grant R-82811201 to L.H. and grant 4953 to C.Z.). The manuscript contents are solely the responsibility of the grantee and do not necessarily represent the official views of the EPA. Further, the EPA does not endorse the purchase of any commercial products or services mentioned in the publication. Research described in this article was conducted under contract to the HEI, an organization jointly funded by the EPA and certain motor vehicle and engine manufacturers. The contents of this article do not necessarily reflect the views of HEI, or its sponsors, nor do they necessarily reflect the views and policies of the EPA or motor vehicle and engine manufacturers. Author contributions: Conceptualization: L.H., C.Z.; Funding acquisition: C.Z.; Investigation: L.H., I.D.; Methodology: L.H., C.C., I.D., C.Z., F.D.; Project administration: F.D., C.Z.; Supervision: C.Z.; Visualization: L.H., J.R., C.Z.; Writing – original draft: L.H., C.Z.; Writing – review and editing: L.H., C.C., I.D., J.R., F.D., C.Z. Competing interests: The authors declare no competing interests. Data and materials availability: The online data exploration tool is available online (47). Code to support analyses in this work is documented on GitHub (48). Coal PM2.5 source impacts and coal EGU information are available for download from the Open Science Framework (49). CMS Henneman et al., Science 382, 941–946 (2023) 24 November 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E Medicare datasets, in accordance with the data use agreement, must be acquired from the Centers for Medicare & Medicaid Services (28). Instructions for acquiring CMS data and code for processing this data are available on GitHub (50). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse. In the interest of rapid dissemination of results with immediate public health relevance, the author will make the Author Accepted Manuscript (AAM) version available under a CC BY public copyright license. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf4915 Materials and Methods Supplementary Text Figs. S1 to S10 Tables S1 and S2 References (51–64) MDAR Reproducibility Checklist Submitted 28 October 2022; accepted 2 October 2023 10.1126/science.adf4915 Henneman et al., Science 382, 941–946 (2023) 24 November 2023 6 of 6
10.1126_science.adf4197
RES EARCH STRUCTURAL BIOLOGY Reorientation of INO80 on hexasomes reveals basis for mechanistic versatility Hao Wu1†, Elise N. Muñoz1,2†, Laura J. Hsieh1, Un Seng Chio1, Muryam A. Gourdet1,2*, Geeta J. Narlikar1*, Yifan Cheng1,3* Unlike other chromatin remodelers, INO80 preferentially mobilizes hexasomes, which can form during transcription. Why INO80 prefers hexasomes over nucleosomes remains unclear. Here, we report structures of Saccharomyces cerevisiae INO80 bound to a hexasome or a nucleosome. INO80 binds the two substrates in substantially different orientations. On a hexasome, INO80 places its ATPase subunit, Ino80, at superhelical location –2 (SHL –2), in contrast to SHL –6 and SHL –7, as previously seen on nucleosomes. Our results suggest that INO80 action on hexasomes resembles action by other remodelers on nucleosomes such that Ino80 is maximally active near SHL –2. The SHL –2 position also plays a critical role for nucleosome remodeling by INO80. Overall, the mechanistic adaptations used by INO80 for preferential hexasome sliding imply that subnucleosomal particles play considerable regulatory roles. I n eukaryotes, central nuclear processes such as gene expression, DNA replication, and DNA repair are coordinated with dy- namic changes in chromatin states (1–3). ATP-dependent chromatin-remodeling en- zymes play essential roles in catalyzing such changes. These enzymes are broadly catego- rized into four major families: SWI/SNF, ISWI, CHD, and INO80 (4, 5). Each of these enzymes contains a core remodeling ATPase subunit and several auxiliary subunits that regulate the core ATPase. It has typically been presumed that the preferred substrate of these enzymes is a nucleosome, the smallest unit of chromatin containing ~147 base pairs (bp) of DNA wrapped around an octamer of histone proteins (6). Con- sistent with this assumption, between them, these four classes slide the histone octamer, exchange histone variants, and transfer en- tire octamers (5, 7). The INO80 complex has been shown to play roles in regulating transcription, DNA repli- cation, and DNA repair (8–11). However, how INO80’s biochemical activities relate to its diverse biological roles is not well understood. Unlike remodelers from other families, in which the core ATPase subunits bind the nu- cleosome near superhelical location 2 (SHL +2 or SHL −2), Ino80, the core ATPase subunit of the INO80 complex, binds nucleosomes near SHL –6 or SHL −7 (fig. S1A) (12–14). It has been speculated that this key difference in nucleo- some engagement reflects a fundamentally dif- ferent remodeling mechanism (15, 16). Indeed, we showed that the preferred substrate of the 1Department of Biochemistry and Biophysics, University of California San Francisco, San Francisco, CA 94158, USA. 2Tetrad Graduate Program, University of California San Francisco, San Francisco, CA 94158, USA. 3Howard Hughes Medical Institute, University of California San Francisco, San Francisco, CA 94158, USA. *Corresponding author. Email: muryam.gourdet.mg@gmail.com (M.A.G.); geeta.narlikar@ucsf.edu (G.J.N.); yifan.cheng@ucsf.edu (Y.C.) †These authors contributed equally to this work. Wu et al., Science 381, 319–324 (2023) 21 July 2023 Saccharomyces cerevisiae INO80 complex is not a nucleosome but a hexasome, which is a subnucleosomal particle that lacks a histone H2A-H2B dimer (17). Hexasomes are gener- ated during transcription and may also be formed during DNA replication and repair (18–21). Further, INO80’s activity on nucleo- somes is more dependent on flanking DNA length than on hexasomes (17, 22). These re- sults suggested that INO80 has the versatility to act on hexasomes or nucleosomes based on the density of nucleosomes and hexasomes at a given locus. However, fundamental mecha- nistic questions remain. It is not clear how INO80 can act on both nucleosomes and hex- asomes, which differ substantially in their structures. It is also unclear why INO80 has different flanking DNA length dependencies on hexasomes versus nucleosomes. Here, we report cryo–electron microscopy (cryo-EM) structures of endogenously purified S. cerevisiae INO80 bound to a hexasome and a nucleosome. We found that INO80 binds hexasomes and nucleosomes in opposite ori- entations, with Ino80 binding near SHL –2 on hexasomes and near SHL –6 or SHL −7 on nucleosomes. The location of the Arp8 module suggests how flanking DNA length differen- tially regulates nucleosome and hexasome slid- ing. DNA gaps near SHL –2 inhibit sliding of both substrates by INO80. Our findings pro- vide mechanistic insights into how INO80 slides both hexasomes and nucleosomes. Structures of the INO80-hexasome and INO80-nucleosome complexes To visualize how INO80 binds to a hexasome or a nucleosome, we prepared hexasomes and nucleosomes on the same DNA templates containing the 147-bp, 601-nucleosome posi- tioning sequence with 80 bp of additional DNA as described previously (+80H and +80N; with definition explained in Fig. 1A; fig. S1, A and B; and the supplementary text) (17, 23, 24). Complexes were formed by mix- ing hexasomes or nucleosomes with endoge- nously purified S. cerevisiae INO80 without adding nucleotide (fig. S1, C to H). We determined cryo-EM structures of the INO80-hexasome complex in three different conformational snapshots (Fig. 1, B and C, and figs. S2 to S6). The overall shape of INO80 is similar within these structures and also to previously published structures of the nucleosome in complex with human (12) and Chaetomium thermophilum (14) INO80. Using prior convention, we grouped subunits of the INO80 complex into four modules: the Rvb mod- ule (Rvb1/Rvb2), the Arp8 module (Arp8/Arp4/ Actin/Ies4 and Taf14), the Ino80 module (Ino80/ Ies2), and the Arp5 module (Arp5/Ies6). The Ino80 protein consists of three major regions: the N- terminal domain, the HSA region (Ino80HSA), and the ATPase domain (Ino80ATPase). Detailed descriptions of these modules in our structures are provided in the supplementary text. Although the INO80 architecture appears similar to that in the INO80-nucleosome struc- tures, a major difference is that it is rotated ~180° on a hexasome compared with a nu- cleosome (Fig. 1, B to E). We identified two primary interactions between INO80 and the hexasome: Ino80ATPase binds the hexasome near SHL –3 (class 1), SHL –2.5 (class 2), or SHL –2 (class 3), and the Arp5/Ies6 module binds near SHL +1, SHL +1.5, or SHL +2 (fig. S6, A and B), respectively. Class 3 is the predom- inant INO80-hexasome class. All Ino80ATPase locations on hexasomes are different than those on nucleosomes, which are near SHL –6 or SHL –7 (12–14). However, the Ino80 orien- tation on hexasomes is consistent with struc- tures of other major chromatin remodelers on nucleosomes such as S. cerevisiae ISW1 (25–27), Chd1 (28–30), RSC (31–33), Snf2 (34), and in particular the SWR1 complex (35), which is from the same subfamily as the INO80 complex. In these structures, the ATPase do- mains interact with nucleosomes near either SHL +2 or SHL –2 (Fig. 1E). Loss of an H2A-H2B dimer in a hexasome causes an additional ~35 bp of DNA to un- wrap from the histone core (free DNA) (Fig. 1A and fig. S1B). Comparison of our hexa- some structures with an unbound hexasome (PDB: 6ZHY) (36) reveals different levels of further DNA unwrapping. In class 1, the hex- asome is almost identical to the unbound hexasome, without detectable additional DNA unwrapping. The level of DNA unwrapping increases as the Ino80ATPase-binding position changes from SHL –3 (class 1) to SHL –2 (class 3) (Fig. 2 and fig. S6C). For comparison, we also determined structures of S. cerevisiae INO80 bound to a nucleosome and captured two conformational snapshots (class 1 and 2) from the same dataset (figs. S7 to S9 and supplementary text). Ino80ATPase 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Structure of the INO80-hexasome complex reveals large rotation. (A) Cartoon illustration of a +X nucleosome and a +X hexasome. H2A-H2B dimer proximal to the flanking DNA (entry side dimer) is shown in cyan; H3-H4, light gray; 601 DNA, dark gray; flanking DNA, orange; additional free (unwrapped) DNA is also shown in cyan; SHLs are shown as yellow dots; DNA from the bottom gyre is shown as a dotted line. (B) Two different views of cryo-EM density map of the INO80-hexasome complex (class 3). (C) Atomic model of the INO80- hexasome complex (class 3) viewed in the same orientation as the map is viewed in (B). (D) Cryo-EM density map of the C. thermophilum INO80-nucleosome complex [EMDB: 4277 (14)] displayed with its nucleosome dyad and H3-H4 tetramer aligned with that of the hexasome in the right panel of (B). Note that INO80 on a hexasome rotates ~180° from where it sits on a nucleosome when keeping the nucleosome-hexasome dyad and H3-H4 aligned. (E) Structural comparisons of INO80-nucleosome complex (left), the SWR-nucleosome complex (middle), and the INO80-hexasome complex (right), with the dyad and H3-H4 of nucleosome and hexasome aligned. Wu et al., Science 381, 319–324 (2023) 21 July 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Conformational snapshots of INO80-hexasome complexes. Comparison of DNA from each INO80-hexasome class (blue) with DNA from an unbound hexasome (PDB: 6ZHY, gray), showing the degree of DNA unwrapping (top row) and binding locations of Ino80ATPase and Arp5 (bottom row). in class 1 is located near SHL –7, similar to its location in the human INO80-nucleosome structure (12), whereas in class 2, it binds near SHL –6, similar to the C. thermophilum struc- ture (14) (fig. S9, A and C). The Arp5/Ies6 module interacts with the nucleosome near SHL –3 and SHL –2 (fig. S9D), respectively. These observations are also consistent with previous findings showing that nucleosomal DNA between SHL –5 and SHL –7 is protected by INO80 (13). The SHL –2 position plays a critical role in nucleosome and hexasome sliding We observed that Ino80ATPase engages the hex- asome predominantly near SHL –2. These results raise the possibility that Ino80ATPase acts near SHL –2 when sliding hexasomes. By contrast, consistent with prior findings (12, 14), we observed that Ino80ATPase engages the nucleosome near two positions, SHL –7 and SHL –6. Also as previously proposed, our findings are consistent with the possibility that Ino80ATPase acts near SHL –6 when sliding nucleosomes (13). A commonly used assay to identify the DNA location from which the ATPase domain of a remodeler acts to trans- Wu et al., Science 381, 319–324 (2023) 21 July 2023 locate DNA is to place a single nucleotide gap at the proposed site of action and test whether the gap inhibits DNA translocation (37–39). Therefore, to directly test the importance of the SHL –6 and SHL –2 locations, we as- sembled nucleosomes and hexasomes with single base gaps near SHL –2 or SHL –6 and measured INO80 activity using a gel-based sliding assay (Fig. 3A). We found that a gap at SHL –6 inhibited INO80’s sliding activity on nucleosomes by ~200-fold, but so did a gap at SHL –2 (Fig. 3, B to G). By contrast, a gap at SHL –6 did not inhibit INO80’s sliding activity on hexasomes, but a gap at SHL –2 slowed hexasomes slid- ing by ~2000-fold (Fig. 3, B to G). These results are consistent with Ino80ATPase acting near SHL –2 when sliding hexasomes and raise new questions about why both the SHL –2 and SHL –6 locations are critical for nucleo- some sliding by INO80. We describe possible explanations in the Discussion. Role of the Arp8 module in flanking DNA length dependence S. cerevisiae INO80 slides +40 nucleosomes ~100-fold more slowly than +80 nucleosomes (17, 22). However, sliding hexasomes is less flanking DNA dependent. Our structures sug- gest that the Arp8 module requires ~40 bp of DNA for appropriate engagement. In class 1 of the INO80-hexasome structure, Arp8 en- gages with the ~35 bp of DNA unwrapped from removal of the H2A-H2B dimer and an additional ~5 bp of flanking DNA. In class 3 of the INO80-hexasome structure, the Arp8 module engages entirely with ~40 bp of un- wrapped DNA that now includes additional DNA unwrapped relative to the unbound hex- asome (Fig. 4). Conversely, in class 2 of the INO80-nucleosome structure, the Arp8 mod- ule engages entirely with flanking DNA, which is consistent with previous findings (40) (Fig. 4). Our structural data with hexasomes, along with the previous data with nucleosomes, suggest that 40 bp may be the minimum amount of DNA needed for the Arp8 module to bind and that proper Arp8 module engage- ment is essential for maximal remodeling activity (40). Altered interactions by the Arp5 module To understand why Ino80 may not bind a nu- cleosome directly near SHL –2, we compared 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Inhibition of DNA translocation at specific SHL sites influences nucleosome and hexasome sliding by INO80. (A) Cartoon illustration of a +80 nucleosome (left) and a +80 hexasome (right) with approximate locations of site-specific single base gaps indicated. Colors are the same as in Fig. 1A. (B and C) Example gels and time courses of native gel–based remodeling assays of WT INO80 on +80 nucleosomes with no gap, a gap near SHL –2, and a gap near SHL –6. (D and E) Example gels and time courses of native gel–based remodeling assays of wild-type INO80 on +80 hexasomes with no gap, a gap near SHL –2, and a gap near SHL –6. (F and G) Average observed rate constants of INO80 sliding activity. kobs (min−1): +80N: 1.551 ± 0.1846; +80N Gap @ SHL –2: 0.005995 ± 0.001054; +80N Gap @ SHL –6: 0.006497 ± 0.0007117; +80H: 1.01 ± 0.1668; +80H Gap @ SHL –2: 0.000379 ± 0.0002849; +80H Gap @ SHL –6: 1.213 ± 0.2209. Data represent the mean ± SEM for three technical replicates performed under single-turnover conditions with saturating enzyme and ATP. interactions made by Arp5/Ies6 in hexasomes versus nucleosomes (see the supplementary text). When INO80 binds to a hexasome, the Arp5/Ies6 regions used in the context of a nu- cleosome are repurposed for different interac- tions. Modeling the missing H2A-H2B dimer into our INO80-hexasome structure reveals steric clashes of the Arp5 module with the en- try side proximal H2A-H2B dimer and with part of the DNA that wraps around the H2A- H2B dimer (fig. S11). These clashes could be avoided if the H2A-H2B dimer is suffi- ciently dislodged. To test for this possibility, we inhibited dimer dislodgement by intro- ducing a site-specific disulfide cross-link be- tween the two H2A molecules (N38C) (41) or promoted dimer dislodgement by using an H2A mutant (R81A) that destabilizes the H2A- H2B/H3-H4 interface (42) (fig. S12, A, B, and H). The disulfide cross-link did not inhibit nu- cleosome sliding, and the H2A mutant did not promote nucleosome sliding (fig. S12, C to G), indicating that complete dimer dislodgement is not necessary for INO80-mediated nucleo- some sliding. In the absence of dimer dislodge- ment, another way to avoid these clashes could be by substantial rearrangement of the Arp5 module together with subtle rearrangements of the H2A-H2B dimer (fig. S9E). Discussion Implications of the INO80-hexasome structure for nucleosome sliding by INO80 The major conformation of the INO80-hexasome complex (class 3) has Ino80ATPase near SHL –2 and ~15 bp of unwrapped DNA from the entry site in addition to the ~35 bp of DNA that is unwrapped from removal of an H2A- H2B dimer. The placement of Ino80ATPase near SHL –2 is consistent with how the ATPase subunits of other remodelers bind the nu- cleosome. Together with our prior finding that hexasomes are remodeled faster than nucleo- somes, these results suggest that the class 3 structure represents the sliding-competent conformation of INO80 on hexasomes (Fig. 5A and fig. S13A). By contrast, the states of INO80 bound to a nucleosome have Ino80ATPase bound near either SHL –6 or SHL –7, also consistent with previous findings. These differences raise the question of whether the INO80-nucleosome structures represent sliding-competent confor- mations or if a rearrangement of Ino80ATPase to SHL –2 is necessary to achieve efficient nucleo- some sliding. Previous cross-linking studies have shown that detachment of nucleosomal DNA from Wu et al., Science 381, 319–324 (2023) 21 July 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. The Arp8 module engages different regions of DNA in nucleosomes versus hexasomes. Overlay of atomic models of the hexasome (class 1 and class 3) and the nucleosome (class 2) with the Arp8 module (PDB: 8A5O), aligned by the H3-H4 tetramer. Fig. 5. Model of INO80-induced hexasome and nucleosome sliding. (A) Hexasome sliding. The Ino80ATPase samples different positions between SHL –3 and SHL –2 but binds predominantly near SHL –2. The INO80 complex becomes sliding competent when Ino80ATPase engages near SHL –2. (B) Nucleosome sliding. INO80 initially binds with Ino80ATPase at SHL –7 or SHL –6. Upon ATP hydrolysis, Ino80ATPase moves toward SHL –2, where INO80 becomes sliding competent. H2A-H2B close to the entry site occurs during INO80 remodeling (13). Our data show that progressively more DNA is unwrapped as Ino80ATPase binds closer to SHL –2 on hex- asomes (Fig. 2 and fig. S6C). Together, these results suggest that DNA unwrapping is coupled to Ino80ATPase accessing its most sliding-competent state. Footprinting studies have shown that whereas binding of INO80 to nucleosomes mainly protects nucleosomal DNA from SHL –5 to SHL –7 and near SHL –3, there is modest but detectable protection near SHL –2 (13). Nicks and gaps between SHL –7 and SHL –2 have been shown to inhibit nucleosome sliding to different extents (13, 43). Here, we show that site-specific gaps near SHL –2 or SHL –6 substantially inhibit INO80’s sliding of nucleosomes (by ~200 fold). DNA gaps are commonly used to identify the site of action of the ATPase domain of remodelers (37–39). We therefore speculate that INO80 initially binds the nucleosome with Ino80ATPase near SHL –6 or SHL –7, and that this is fol- lowed by an ATP-dependent rotation around the nucleosome to position Ino80ATPase near SHL –2, from which Ino80ATPase then trans- locates nucleosomal DNA (Fig. 5B and fig. S14A). A gap near SHL –6 would then inhibit ATP-dependent movement of Ino80ATPase on the nucleosome, whereas the gap near SHL –2 would inhibit translocation of nucleosomal DNA by INO80 relative to the histone octa- mer (fig. S14). Single-molecule fluorescence resonance energy transfer studies have iden- tified an ATP-dependent pause phase be- fore ATP-dependent nucleosome sliding (22). The pause could represent the reorientation of Ino80ATPase from SHL –6 or SHL –7 toward SHL –2 and add a step that slows remodeling of nucleosomes compared with hexasomes. Simply placing the INO80 complex as is on nucleosomes with the Ino80ATPase near SHL –2 results in steric clashes of the Arp5 module with the nucleosome (fig. S11). Although partial H2A-H2B dimer dislodgment, as previously proposed (17), could avoid such clashes, our biochemical data here indicate that dimer dis- lodgement is not essential for nucleosome sliding by INO80 (fig. S12). More structural studies are needed to understand how INO80 might rotate around a nucleosome. Alternatively, a gap near SHL –2 may affect the action of the Arp5 module. For such a scenario, we speculate that Ino80ATPase trans- locates DNA near SHL –6, and effective trans- location also requires action of the Arp5 module near SHL –2, as previously proposed (12, 14). A gap near SHL –6 would then in- hibit translocation of nucleosomal DNA by Ino80ATPase, and a gap near SHL –2 would inhibit productive engagement of the Arp5 module (fig. S15). Clearly distinguishing between these two mod- els will require substantial additional structural Wu et al., Science 381, 319–324 (2023) 21 July 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E analysis of INO80-remodeling intermediates on nucleosomes. Implications for hexasome sliding by INO80 Our structures provide a view into how INO80 engages a hexasome. In the predominant INO80-hexasome structure, Ino80ATPase binds near SHL –2. A site-specific gap near SHL –2 substantially inhibits INO80’s sliding of hex- asomes (~2000 fold), whereas a gap near SHL –6 does not have a major effect. We therefore hypothesize that Ino80ATPase bound at SHL –2 on a hexasome represents the active structure. Compared with the subtle changes at SHL –2 observed when other remodelers bind nucleosomes (16), the additional 15 bp of unwrapped DNA (up to SHL –2.5) in class 3 substantially loosens histone DNA interactions and thus may allow more ready translocation from SHL –2. We further propose that the new contacts made by the Arp5/Ies6 module with the exposed H3-H4 surface provide an anchor that allows the Ino80 motor to efficiently pump DNA through the hexasome. These findings also explain the differential effects of the Arp5/Ies6 module on hexasome versus nucleosome sliding (17). The location of the Arp8 module is also different on hexasomes than on nucleosomes. On nucleosomes the Arp8 module binds ~40 bp entirely on the flanking DNA (Fig. 4). In the most prevalent INO80-hexasome state (class 3), the Arp8 module is bound entirely to the unwrapped DNA, substantially reducing the need to bind flanking DNA (Fig. 4). These different binding modes of the Arp8 module could explain why hexasome sliding by INO80 is less dependent on flanking DNA length compared with nu- cleosome sliding. RE FE RENCES AND N OT ES 1. R. Bar-Ziv, Y. Voichek, N. Barkai, Genome Res. 26, 1245–1256 (2016). 2. A. E. Ehrenhofer-Murray, Eur. J. Biochem. 271, 2335–2349 (2004). 3. M. R. Hübner, D. L. Spector, Annu. Rev. Biophys. 39, 471–489 (2010). 4. C. R. Clapier, B. R. Cairns, Annu. Rev. Biochem. 78, 273–304 (2009). 5. C. Y. Zhou, S. L. Johnson, N. I. Gamarra, G. J. Narlikar, Annu. Rev. Biophys. 45, 153–181 (2016). 6. K. Luger, A. W. Mäder, R. K. Richmond, D. F. Sargent, T. J. Richmond, Nature 389, 251–260 (1997). 7. C. R. Clapier, J. Iwasa, B. R. Cairns, C. L. Peterson, Nat. Rev. Mol. Cell Biol. 18, 407–422 (2017). 8. Y. Bao, X. Shen, Mutat. Res. 618, 18–29 (2007). 9. A. J. Morrison, X. Shen, Nat. Rev. Mol. Cell Biol. 10, 373–384 (2009). 10. J. Poli, S. M. Gasser, M. Papamichos-Chronakis, Philos. Trans. R. Soc. Lond. B Biol. Sci. 372, 20160290 (2017). 11. X. Shen, G. Mizuguchi, A. Hamiche, C. Wu, Nature 406, 541–544 (2000). 12. R. Ayala et al., Nature 556, 391–395 (2018). 13. S. Brahma et al., Nat. Commun. 8, 15616 (2017). 14. S. Eustermann et al., Nature 556, 386–390 (2018). 15. J. Markert, K. Luger, Trends Biochem. Sci. 46, 41–50 (2021). 16. L. Yan, Z. Chen, Trends Biochem. Sci. 45, 217–227 (2020). 17. L. J. Hsieh et al., Mol. Cell 82, 2098–2112.e4 (2022). 18. S. Henikoff, Cold Spring Harb. Perspect. Med. 6, a026666 (2016). 19. M. L. Kireeva et al., Mol. Cell 9, 541–552 (2002). 20. O. I. Kulaeva, F. K. Hsieh, V. M. Studitsky, Proc. Natl. Acad. Sci. U.S.A. 107, 11325–11330 (2010). 21. S. Ramachandran, K. Ahmad, S. Henikoff, Mol. Cell 68, 1038–1053.e4 (2017). 22. C. Y. Zhou et al., Mol. Cell 69, 677–688.e9 (2018). 23. R. F. Levendosky, G. D. Bowman, eLife 8, e45472 (2019). 24. R. F. Levendosky, A. Sabantsev, S. Deindl, G. D. Bowman, eLife 5, e21356 (2016). 25. J. P. Armache et al., eLife 8, e46057 (2019). 26. S. Chittori, J. Hong, Y. Bai, S. Subramaniam, Nucleic Acids Res. 47, 9400–9409 (2019). 27. L. Yan, H. Wu, X. Li, N. Gao, Z. Chen, Nat. Struct. Mol. Biol. 26, 258–266 (2019). 28. L. Farnung, S. M. Vos, C. Wigge, P. Cramer, Nature 550, 539–542 (2017). 29. I. M. Nodelman et al., Nat. Struct. Mol. Biol. 29, 121–129 (2022). 30. R. Sundaramoorthy et al., eLife 7, e35720 (2018). 31. A. B. Patel et al., eLife 8, e54449 (2019). 32. F. R. Wagner et al., Nature 579, 448–451 (2020). 33. Y. Ye et al., Science 366, 838–843 (2019). 34. X. Liu, M. Li, X. Xia, X. Li, Z. Chen, Nature 544, 440–445 (2017). 35. O. Willhoft et al., Science 362, eaat7716 (2018). 36. L. C. Lehmann et al., Cell Rep. 33, 108529 (2020). 37. A. Saha, J. Wittmeyer, B. R. Cairns, Nat. Struct. Mol. Biol. 12, 747–755 (2005). 38. R. Schwanbeck, H. Xiao, C. Wu, J. Biol. Chem. 279, 39933–39941 (2004). 39. M. Zofall, J. Persinger, S. R. Kassabov, B. Bartholomew, Nat. Struct. Mol. Biol. 13, 339–346 (2006). 40. F. Kunert et al., Sci. Adv. 8, eadd3189 (2022). 41. T. D. Frouws, P. D. Barth, T. J. Richmond, J. Mol. Biol. 430, 45–57 (2018). AC KNOWLED GME NTS We thank L. Zheng for advice on model building, Z. Yu for providing GO grids, J. Tretyakova for expressing and purifying histones, U. Kaur for providing INO80 WT enzyme, members of the Narlikar and Cheng laboratories for helpful discussions, C. Wu and A. Ranjan for sharing unpublished data showing inhibition of INO80 activity on nucleosomes containing a gap near SHL –2, and S. Diendl and A. Sabantsev for detailed advice on generating DNAs with site-specific base gaps. The cryo-EM facility at UCSF is managed by D. Bulkley and G. Gilbert; computation at the Cheng laboratory is supported by M. Harrington and J. Li. Funding: This work was supported by the National Institutes of Health (NIH grant R35GM140847 to Y.C., grant R35GM127020 to G.J.N., grant F31GM136187 to M.A.G., grant F31GM142271 to E.N.M., and grant F32GM137463 to U.S.C.) and by the American Cancer Society (Roaring Fork Valley Research Fund Postdoctoral Fellowship PF-18-155- 01-DMC to L.J.H.). Equipment at the UCSF cryo-EM facility was partially supported by the NIH (grants S10OD020054, S10OD021741, and S10OD025881). Y.C. is an investigator at the Howard Hughes Medical Institute. Author contributions: H.W., M.A.G., L.J.H., and E.N.M. purified INO80; M.A.G., L.J.H., and E.N.M. assembled and purified hexasomes and nucleosomes; H.W. performed cryo-EM; E.N.M. performed and quantified all biochemical experiments; U.S.C. generated the H2A R81A mutant; M.A.G., G.J.N., and Y.C. conceived and oversaw the project; and all authors participated in interpretation and discussion of the results and writing of the manuscript. Competing interests: Y.C. is scientific advisory board member of ShuiMu BioSciences Ltd. Data and materials availability: For the core INO80 of the INO80-Hexasome complex (class 1, class 2, and class 3) and the core INO80 of the INO80-Nucleosome complex (class 1 and class 2), the coordinates are deposited in the Protein Data Bank with the accession codes 8ETS, 8ETU, 8ETW, 8EU9, and 8EUF; the cryo-EM density maps are deposited in the Electron Microscopy Data Bank (EMDB) with the accession codes EMD-28597, EMD-28599, EMD-28601, EMD-28609, and EMD-28613. For the hexasome of the INO80-Hexasome complex (class 1, class 2, and class 3) and the nucleosome of the INO80-Nucleosome complex (class 1 and class 2), the coordinates are deposited in the Protein Data Bank with the accession codes 8ETT, 8ETV, 8EU2, 8EUE, and 8EUJ; the cryo-EM density maps are deposited in the Electron Microscopy Data Bank (EMDB) with the accession codes EMD-28598, EMD-28600, EMD-28602, EMD-28612, and EMD-28614. License information: This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the author- accepted manuscript (AAM) of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf4197 Materials and Methods Supplementary Text Figs. S1 to S15 Tables S1 and S2 References (44–71) MDAR Reproducibility Checklist 42. K. Lehmann et al., Sci. Rep. 7, 13303 (2017). 43. F. Mueller-Planitz, H. Klinker, P. B. Becker, Nat. Struct. Mol. Biol. 20, 1026–1032 (2013). Submitted 21 October 2022; accepted 17 June 2023 Published online 29 June 2023 10.1126/science.adf4197 Wu et al., Science 381, 319–324 (2023) 21 July 2023 6 of 6
10.1126_science.adf5273
RES EARCH MECHANOCHEMISTRY Acceleration of Diels-Alder reactions by mechanical distortion Yerzhan S. Zholdassov1,2,3, Li Yuan4, Sergio Romero Garcia5, Ryan W. Kwok2,3, Alejandro Boscoboinik4, Daniel J. Valles1,2,3, Mateusz Marianski2,3, Ashlie Martini5, Robert W. Carpick4, Adam B. Braunschweig1,2,3* Challenges in quantifying how force affects bond formation have hindered the widespread adoption of mechanochemistry. We used parallel tip-based methods to determine reaction rates, activation energies, and activation volumes of force-accelerated [4+2] Diels-Alder cycloadditions between surface-immobilized anthracene and four dienophiles that differ in electronic and steric demand. The rate dependences on pressure were unexpectedly strong, and substantial differences were observed between the dienophiles. Multiscale modeling demonstrated that in proximity to a surface, mechanochemical trajectories ensued that were distinct from those observed solvothermally or under hydrostatic pressure. These results provide a framework for anticipating how experimental geometry, molecular confinement, and directed force contribute to mechanochemical kinetics. T he great majority of organic reactions carried out in industry and research laboratories rely on strictly solvothermal activation, in which solvent and heat act in concert to drive reactions along a trajectory toward products (1). The organic solvents used in these reactions account for >60% of all chemical manufacturing waste (2) and are often toxic (3), while as a result of the energy demand of solvothermal syntheses, the chemical industry consumes 37% of the total energy used in manufacturing (4). As such, there is a critical need to supplement and ideally replace solvothermal chemistry with less wasteful activation approaches. Mechani- cally activated organic chemistry (5)—in which force, rather than heat and solvent or hydro- static pressure, drives the making and break- ing of covalent bonds—can help to ameliorate the waste and energy challenges of organic synthesis because the reactions are run neat or with minimal solvent (6), and energy can be provided more efficiently through mechanical rather than thermal means (7). Even when solvents are added to mechanochemical reac- tions, in a process called liquid-assisted grind- ing, they are added in comparatively small amounts (8). Another benefit is that mecha- nochemically activated conditions often pro- duce stereoisomers and regioisomers that are not favored, or cannot be produced at all, under solvothermal conditions (9), although why such mechanically favored products arise is often not well understood. 1The Advanced Science Research Center, Graduate Center of the City University of New York, New York, NY 10031, USA. 2Department of Chemistry, Hunter College, New York, NY 10065, USA. 3Ph.D. Program in Chemistry, Graduate Center of the City University of New York, New York, NY 10016, USA. 4Department of Mechanical Engineering and Applied Mechanics, University of Pennsylvania, Philadelphia, PA 19104, USA. 5Department of Mechanical Engineering, University of California, Merced, CA 95343, USA. *Corresponding author. Email: abraunschweig@gc.cuny.edu Despite its many potential advantages, organ- ic mechanochemistry has not been widely adopted in chemical synthesis and manu- facturing. This can be attributed in part to substantial remaining gaps in the understand- ing of reaction kinetics under mechanochemical activation. Until physical models are devel- oped that reliably account for experimental observations, the products of mechanochemical reactions cannot be predicted, and the reac- tions cannot be easily translated to large scales. Mechanochemical reactions are typi- cally carried out in ball mill (10) or planetary mill reactors (11) or in twin-screw extruders (12), for which it is difficult to determine even the most basic reaction parameters, such as the magnitude of the applied force (F), the reaction time under force (t), or conversion at different time points. Polymer mechano- chemistry (13–15) entails the use of mechan- ical energy to activate mechanophores (16–18) embedded within long polymer chains: Soni- cation (19, 20), atomic force microscopy (AFM) (21, 22), and/or laser pulses (23) apply me- chanical energy that ruptures the mechani- cally susceptible bonds in the mechanophores. However, these methods are used to initiate bond rupture, and as a consequence, such studies do not explain bond-formation events in mills and extruders. This difficulty in acquir- ing quantitative experimental data on bond formation under stress complicates the develop- ment and validation of generalizable kinetic models of mechanochemical bond-forming reactions. Furthermore, because mechano- chemical reactions are carried out on powders, rather than well-solvated molecules, the kine- tics of grinding to expose molecular reactants adds another layer of complexity to kinetic models. For example, the rate of pericyclic [4+2] Diels-Alder cycloadditions in ball (10) and planetary (24) mills are dependent on particle size (25), shaking frequency (10), ball mass (25), and vessel loading (26) in addition to the molecular-scale factors that are known to affect rates, such as reactant structure and temperature (T) (27). Given all these macroscopic factors, it has been challenging to explain how mechanical energy acts on the reaction potential energy surface. One such explanation is the “hotspot” model (28), which postulates that localized heating occurs when milling balls collide with surfaces and with each other, causing the re- lease of kinetic energy. Rate increases in mechanochemical reactions are thus posited to arise because of conventional thermal acti- vation at the collision sites. Although some local heating is observed upon milling, these temperature increases are small and are not commensurate with observed reaction rate increases (29). The other dominant kinetic theory of mechanochemistry argues that grinding increases the surface area of powders, increasing the collision probability of reac- tants, which explains the dependence of rates on powder size (25). Ultimately, current kinetic models of organic mechanochemistry—the molecular-scale hotspot theory and the macro- scopic model that considers increases in area during grinding—are incomplete because they fail to account for all experimental observations. In particular, they fail to account for the forma- tion of different isomers under mechano- chemical activation versus strictly solvothermal activation, which indicates that F can move reactants along a different reaction trajectory than that from heat. As such, any accurate molecular-scale kinetic model of mechano- chemical reactions must consider how F distinctly alters reaction trajectories and explain quantitatively how F affects reaction barriers and, in turn, reaction rates. With the physical understanding inherent to an accu- rate molecular-scale model of mechanical reactivity, the reactions that are susceptible to mechanical activation could be anticipated and their rates and products predicted accu- rately, leading to the wider adoption of sustain- able mechanochemical methods in organic synthesis and chemical manufacturing. We experimentally and computationally investigated the reaction kinetics of mecha- nically activated [4+2] Diels-Alder cycloaddition reactions between dienophiles and surface- confined diene monolayers to measure how force affects reaction rates. Pericyclic reactions were chosen because they proceed in con- certed fashion without intermediates (30), which minimizes challenges in analyzing their kinetics. The reactions were carried out on monolayers so that the complexities associ- ated with grinding powders during milling and reactant availability were not factors in the kinetic analysis, and the effects of F on the free energy of activation (DG‡) and reaction trajectories would be isolated. The experiment Zholdassov et al., Science 380, 1053–1058 (2023) 9 June 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E (Fig. 1) used elastomeric arrays (figs. S7 and S23) that contained 900 pyramidal tips with heights of 21.9 mm and base edge lengths of 31 mm (31, 32) to bring fluorescently labeled dienophiles into contact with monolayers of the tethered diene, anthracene, that was immo- bilized covalently onto the surface of a silica (SiO2) wafer (33, 34). The tip arrays were mounted onto piezoactuators so that they could be pushed into the monolayer with pre- cise control over F and t (32, 35, 36), resulting in a nanoreactor where the anthracene and dienophile react under mechanical activation. The tips themselves were coated with an ink mixture that consisted of a large excess of the fluorescent dienophiles embedded within a polymer matrix that solubilized the dienophiles and ensured that they were accessible to react with the immobilized anthracene (37). We used fluorescence microscopy to track the Fig. 1. Force-activated Diels-Alder reactions on anthracene monolayers. (A) Diels-Alder cyclo- addition reaction between anthracenes immobilized onto SiO2 surfaces and dienophiles under applied force. (B) Structures of dienophiles Alk, Mal, MAcr, and Acr. (C) (i) Elastomeric tip arrays transfer an ink mixture (red coating), consisting of a dienophile and PEG, onto an anthracene-modified surface. The tip arrays are tilted so that different forces are applied by the tips at different locations across the surface. Thick arrows indicate areas of the tip array that exert high force, and thin arrows indicate areas that exert low force. (ii) Upon contact with the surface, the tips form nanoreactors, (iii) where forces are applied that accelerate the Diels-Alder cycloaddition reactions. (iv) After washing the surface, only covalently bound molecules remain on the surface. Diels-Alder adduct formation as a function of F and t. We studied the reaction rates of the anthracene monolayer with four different dienophiles—an alkene (Alk), a maleimide (Mal), a methacrylate (MAcr), and an acryl- ate (Acr)—because they differ in the electron demand (38) and the steric environment (39) around the reactive alkene, factors that af- fect the reaction rates of Diels-Alder reactions (27, 30) under strictly solvothermal conditions, which allowed us to examine how mechano- chemical reactivity trends differ from strictly solvothermal trends. We determined the pres- sure (p) of the reaction using a scale below the tips and finite element analysis modeling (FEM). The fluorescence data were fit to a first-order kinetic model to determine acti- vation parameters, including the activation energy (Ea), which is the energy difference between the transition state and the initial state in the absence of stress; DG‡; and the activation volume (DV‡), which is a measure of the sensitivity of a reaction to mechanical activation (40, 41). These data reveal that rate constant (k) increases of a factor of >10× occur at only a few atmospheres of p, and the sensitivity of this experimental approach is such that the effects of subtle differences in dienophile structure on rates could be deter- mined, revealing that mechanochemical re- activity trends are distinct from solvothermal trends. Molecular dynamics (MD) and density functional theory (DFT) modeling of the sur- face and reaction revealed how molecular distortions account for changes in DG‡, providing a generalizable model for mechanochemical reactivity that suggests that the scope of me- chanically susceptible reactions may be substan- tially broader than previously anticipated. Experimental setup For the printing experiment, we modified an approach we have developed for studying mechanochemical surface reactions quanti- tatively (31, 32), in which the reactions occur under a ~1-cm2 elastomeric polymer tip array (42, 43). The inked 900-tip arrays were mounted onto the z-piezoactuator of an atomic force microscope equipped with a specialized mount for the tip arrays and an x,y–tilting stage. Upon moving the arrays in and out of contact with the surface, each tip in the array prints a two- by-five array of 10 features, in which each feature in a two-by-five array is printed with a different tip-surface contact time t that varies from 1 to 600 s. The substrate is intentionally tilted with respect to the tip-array so that the F applied by individual tips varies systemati- cally across one axis of the surface (36). In this experiment, the effect of 30 different values of F, ranging from 1.91 to 89.4 mN (table S1), and 10 different values of t on reaction con- version could all be assessed from a single printed surface. In addition, each two-by-five pattern generated at a given F is repeated 30 times by the tip arrays, resulting in denser data sets that enable higher-fidelity fitting. After printing, the surface was sonicated in ethyl alcohol (EtOH) for 5 min to remove any phy- sisorbed dienophile and polyethylene glycol (PEG) from the substrate, and F, k, Ea, DV‡, and DG‡ values for the reaction between a fluorescent diene and surface-bound anthra- cene could be determined from the normalized fluorescence intensity I (fluorescence of feature/ fluorescence of background) from images taken from a single printed surface. Characterization of patterned surfaces We relied on several complementary analytic- al techniques to confirm the formation of the anthracene monolayer and, subsequently, the formation of the Diels-Alder adduct. Fluores- cence intensity data derived from fluorescence microscopy images (Fig. 2A) confirmed that the immobilization of the fluorophores occurs only where the tips contact the surface, and that I is dependent on both F and t (Fig. 2B). Time-of-flight secondary ion mass spectro- metry (ToF-SIMS) data (Fig. 2C) show that fluorescent features of the two-by-three patterns are chemically distinct from unprinted areas, with the spectra of the former containing fea- tures consistent with the presence of the dienophile, whereas the latter are consistent with an unreacted anthracene monolayer. X-ray photoelectron spectroscopy (fig. S3), contact angle measurements (fig. S6), and AFM imag- ing (fig. S13) were carried out, and all data are consistent with the proposed modified sur- faces. Control experiments (fig. S12), including patterning of the fluorophores onto a bare SiO2 surface or the patterning of dyes that lack the reactive alkene onto anthracene surfaces, failed to produce fluorescence patterns, fur- ther confirming that the fluorescent patterns are the result of the covalent bond formation that occurs during the Diels-Alder cycloaddition. Although no single surface technique is capa- ble of directly confirming the formation of new covalent bonds in monolayers, the totality of the data makes any explanation of the fluorescent patterns for a reason other than covalent bond formation through a [4+2] Diels-Alder reaction unlikely. Determination of activation parameters To analyze the fluorescence data and ex- tract kinetic parameters, we assumed that the Diels-Alder reaction in this system follows a first-order kinetic rate law (eq. S10) and that fluorescence images can be used to quantify conversion as a function of F and t. We adopted the pseudo–first order assumption because at the interface, the dienophile concentration is substantially greater than the anthracene concentration. We have previously applied this same model to the kinetic data of other Zholdassov et al., Science 380, 1053–1058 (2023) 9 June 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Experimental data for Diels-Alder reactions on anthracene mono- layers with varying force and times. (A) Fluorescence image (lex = 530 to 550 nm and lem = 575 to 750 nm, where lex is the excitation wavelength and lem is the emission wavelength), observed after washing and sonication, of surfaces after patterning with fluorescently labeled dienophile Alk onto the anthracene- modified SiO2 surface. The imaged areas vary from high (top) to low (bottom) F applied to the substrate during the cycloaddition reaction by the tilted tip arrays. Scale bar, 90 mm. (B) Normalized fluorescence intensity (I is fluorescence of feature/fluorescence of background) of the printed features of Alk as a function of F and t. Arrows above the bars indicate the data taken from the fluorescent images in (A); the wide arrows indicate high F, and narrow arrows indicate low F. (C) Time-of-flight secondary-ion mass spectrometry (ToF-SIMS) map (positive ion mode) of a surface patterned with Alk after sonication and washing. (Middle left) Fluorescent image. (Middle right) Corresponding ToF-SIMS map. Scale bar, 100 mm. Mass spectrometer data were taken from unreacted (top) and reacted (bottom) areas of the surface. The peak at mass/charge ratio + fragment corresponding to anthracene, (m/z) = 178.08 is identified as a C14H10 and the peak at m/z = 73.01 is identified as a C4H11N+ fragment corresponding to the secondary dialkyl amine of the rhodamine dye. (D) QCM measurement tracking the formation of the Diels-Alder adduct from the reaction of Mal with anthracene-functionalized QCM crystal in PhMe solution. The abrupt drop in the frequency corresponds to the introduction of new solutions into the QCM. The spike in the frequency and inset correspond to the introduction of a solution of EtOH to wash physisorbed molecules from the surface. (E and F) Plots of surface density of the adduct (Gadd) versus t for the reaction of Alk with the anthracene surface. (G and H) Plots of Gadd versus F for the reaction of Alk with the anthracene surface. Error bars are 1 SD from the mean of three independent measurements. Fits were weighted to the error bars. pericyclic reactions on monolayers under applied F and found good agreement between the model and the experimental data (32). To determine the rate constant k, it was nec- essary to track conversion of anthracenes on the surface to adduct by measuring the grafting density of the anthracene (Ganth) and of the ad- duct (Gadd) in the fluorescent features. A quartz crystal microbalance (QCM) measurement on a SiO2 crystal that was subjected to the anthra- cene monolayer formation protocols provided Ganth,max, the concentration of anthracene in the unreacted monolayer, of 1.12 ± 0.33 mol·nm−2 (fig. S18), which is in good agreement with the value obtained from MD simulations of Ganth,max = 1.20 ± 0.08 mol·nm−2 (figs. S36 and S37). The maximum grafting density of the adduct (Gadd,max) was determined by moni- toring the reaction between an anthracene- modified QCM crystal and a solution of Mal within the fluid cell of the QCM (Fig. 2D). This reaction was selected because it occurs rapidly, even in the absence of F, so that the reaction would proceed to completion in the QCM. The QCM data were consistent with the formation of the adduct with Gadd,max = 0.94 ± 0.14 mol·nm−2, suggesting that only 75 to 84% of the anthracene can react with Mal, at which point the surface is saturated with adduct. For the ensuing calculations, we assumed that the adducts that formed with the differ- ent dienophiles had similar volumes and, in turn, similar Gadd,max. An accurate measure of F at each feature was needed to understand the relationship between reaction rate and F. We determined F using a balance below the printed surface and measuring the recorded weight and feature edge length (fig. S22 and table S1) (35), and these calculations deter- mined that the F across the substrate ranged from 1.91 to 89.4 mN. We conducted FEM of individual pyramidal tips to validate the force measurements, and the relationship between feature length and z-displacement matched the experiments and was nearly indepen- dent of the assumed elastic constants for poly- dimethylsiloxane (PDMS). Furthermore, the values of the simulated F as a function of z-piezo extension could be readily fit to the experimental data by using values of the PDMS elastic constants well within typical measured values (figs. S23 and S24). Last, we considered whether surface roughness could create local- ized asperities of high force and found that average surface roughness (0.41 nm) (fig. S25) was too low to cause any noteworthy effect. With these data in hand, the surface density of the adduct (Gt,F) could be determined from the fluorescence intensity of the features printed at different F and t from It,F using Eq. 1 (cid:4) (cid:1) (cid:3) (cid:5) Gt;F ¼ 1 (cid:2) Imax (cid:2) It;F Imax where Imax is the fluorescence intensity of a feature that has reached saturation with the adduct. This equation assumes Gt,F and It,F (cid:3) Gadd;max ð1Þ Zholdassov et al., Science 380, 1053–1058 (2023) 9 June 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E are linearly related, an assumption that can be used in organic monolayers when the fluo- rophores do not couple (44). The relationship between Gt,F, F, and t for each feature was plotted for all anthracene-dienophile pairs (Fig. 2, E to H, and figs. S27 to S30). In all cases, we observed that Gt,F increased with increasing t and F. Mal is the only dieno- phile that formed adduct in the absence of F, and the rate quickly reached a maximum upon the application of minimal F. This is consistent with known trends for Diels-Alder reactions with normal electron demand, in which the presence of electron withdrawing groups on the dienophile lowers the Ea (27, 30). The plots of Gt,F versus t (Fig. 2, E and F) and Gt,F versus F (Fig. 2, G and H) fit well to a first-order rate model, further validating the pseudo– first order assumption. All fits of kinetic data were weighted to the error bars, and c2 analysis showed that the fits were statistically signifi- cant at a 95% level of confidence. A mathematical framework was needed to derive the activation parameters k, Ea, DV‡, and DG‡ from the experimental data. We adapted the model, developed from examining chemical reactions under hydrostatic pressure (45, 46) and based on the work of Evans and Polanyi (47) and Eyring (48), that considers how p affects k. Under hydrostatic pressure, cyclo- addition reactions have finite and negative DV‡, meaning that p will increase k according to Eq. 2 ln kð Þ ¼ ln Að Þ (cid:2) Ea kBT (cid:2) pDV ‡ kBT ð2Þ where A is the preexponential factor, T is the temperature, and kB is Boltzmann’s constant. This model, which considers the effects of thermal and mechanical activation on the rate, has DG‡ = Ea + pDV‡. We follow the convention, used broadly in the study of the effects of p on chemical kinetics, in which a negative DV‡ acts to reduce DG‡. To apply this model, the mean p in the nanoreactor contacts was determined as a function of F by dividing by the feature area. Our use of mean pressure is consistent with current pub- lished work on probe-based mechanochemistry (32, 49, 50). There was a spatial distribution of that pressure, as seen in our finite element simulations (fig. S23D), but this does not alter the observed trends nor our primary conclu- sion that nonhydrostatic compression increases reaction rates as a result of molecular distor- tion, lowering the reaction energy. With these experimental data, we deter- mined all the activation parameters of interest for each dienophile-anthracene pair. Isothermal first-order Arrhenius plots of lnGanth versus t were fitted, and k at each pressure was deter- mined from the best-fit linear correlation (Fig. 3A and fig. S31) that was weighted to the error bars; errors are reported as the standard error of the fit. For Alk, k increased from 3.8 ± 0.25 × 10−3 s−1 at p = 0.16 MPa to 1.55 ± 0.19 × 10−2 s−1 at p = 0.32 MPa, and this trend of in- creasing k with increasing p is consistent throughout all the data. The greatest change in k was observed for MAcr (from 4.07 ± 0.58 × 10−3 to 1.77 ± 0.77 × 10−1 s−1) and the smallest for Mal (from 4.0 ± 0.050 × 10−2 to 1.57 ± 0.77 × 10−1 s−1). The Ea and DV‡ were determined from the extrapolated y-intercept and slope, respectively, of the linear fit of the plot of lnk versus p (Fig. 3B). The Ea follow the trend MAcr > Acr > Alk > Mal. The mag- nitudes of Ea in the anthracene monolayers, 84.5 to 96.4 kJ·mol−1, are similar in magni- tude to values for Diels-Alder reactions with anthracene reported in the literature of 63.6 to 83.7 kJ·mol−1 (51, 52), and these small differences in magnitude in the Ea between reactions in solution and monolayers can be attributed to differences in molecular structures, reaction geometry, lack of solvent, and the influence of the SiO2 surface. All DV‡ are negative, and the differences in DV‡ between the dienophile- anthracene pairs reflect the relative sus- ceptibilities of their rates to mechanically induced pressure. The DV‡ range from –60.7 × 103 cm3·mol−1 for MAcr to –21.8 × 103 cm3·mol−1 for Mal and follow the identical trend of MAcr > Acr > Alk > Mal, meaning that the reactions with larger Ea are more sensitive to mechanochemical activation. Also, these DV‡ values are substantially greater than the DV‡ of –20 to –45 cm3·mol−1 for Diels-Alder re- actions observed (46) under hydrostatic p in solution, where interfaces do not have a role in the reaction. Last, we calculated the change in free energy of activation DDG‡ = pDV‡ at every p for each dienophile-anthracene pair (Fig. 3C) and found that DDG‡ range from –3.53 to –19.46 kJ·mol−1, with an average value of –8.21 kJ·mol−1, which would cause a ~14-fold rate increase in rate. The relative sensitivity to p shown in Fig. 3B was validated by adding the rhodamine-labeled Mal (Fig. 3B, red) and a MAcr that was labeled with fluorescein (Fig. 3B, green) to the same tip array (fig. S33) and then printing them simultaneously (fig. S34), which led to the fluorescence intensity of the red dye remaining constant as force was in- creased, whereas the intensity of the green dye increased with increasing force (fig. S35). Modeling mechanical distortion We used DFT calculations to generate a theoretical model for the energetics of the reactants and the transition states (TSs) and to understand mechanistically how F acts on the system to lower the reaction energy (DE). Al- though experiments determined DG‡, our calculations and the ensuing theoretical treat- ment involve DE, which is the reaction energy that does not take into account vibrational and entropic effects that are not included in the DFT calculations. Anthracene was simu- lated as attached to a Si4O9H3 cluster to mimic the effect of the surface. DFT calculations predicted the Ea, which is the activation en- ergy for the strictly solvothermal reactions, as defined by Eq. 2, for the Diels-Alder reac- tion to be between 77.1 (Mal) and 90.0 (Alk) kJ·mol−1 (Fig. 4A and fig. S44). These values are within range of the experimental values. The overestimation of the Ea of Alk is likely caused by limiting the reaction system to the anthracene, SiO2 cluster, and dienophile. In doing so, we neglected the destabilizing ef- fects caused by the interactions between the surrounding SiO2 surface and the rhodamine fluorophore. This omission led to interactions Fig. 3. Activation parameters for the Diels-Alder reactions on anthracene monolayers. (A) Isothermal Arrhenius plots of ln(Ganthracene) versus t at different F for the reaction between Alk and the anthracene monolayers. Fits are based on eq. S12. The slope of these lines provide k. (B) Plot of lnk versus p for all four dienophiles, the slope of which provides DV‡, whereas y-intercept provides Ea. Fits were weighted to the error bars, and coefficients of determination (R2) are 0.97 to 0.99. (C) DDG‡ as a function of p. Zholdassov et al., Science 380, 1053–1058 (2023) 9 June 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E in the calculations that stabilized the Alk- anthracene reactant complex, resulting in a higher reaction barrier than was measured experimentally. MD simulations showed that the anthracene responds to tip-induced com- pression by bending toward the surface, that the fraction of molecules with distorted CCH angles increases with increasing pressure (fig. S42), and that compression causes deforma- tion of the molecules themselves (figs. S40 to S42). To model the effect of the mechanical distortion caused by uniaxial stress acting on the reactant from the top, we used DFT to distort the anthracene along three angles— OSiC, SiCC, or CCH (Fig. 4A)—and the geo- metry of the reactant and TS were fully relaxed, while keeping those angles fixed. In this model, we envision that F increases the energy of the reactants by distorting the anthracene in a way that increases its re- activity. The acceleration of the reaction can be then achieved when the distortion energy of the reactants (DEd R) is larger than the dis- tortion energy of the TS (DEd TS) (53). The change in the reaction energy (DDE) that oc- Fig. 4. Mechanical distortion under uniaxial stress. (A) The activation energy Ea and the geometry of the relaxed transition state involving Acr. Blue lines indicate OSiC, SiCC, and CCH angles, distortion of which by the applied F is explored in (B) and (C). All calculations have been performed at M06-2X/6-311+G(d,p) level of theory (31, 32, 57, 58). (B) The relative change (DDE) to the reaction energy (DE) as a function of the CCH angle. The gray box indicates the fully relaxed TS geometry. The bending of the CCH angle leads to the decrease of the DE for MAcr, Acr, and Alk. (C) The effect of distorting the CCH coordinate on the geometry of the TS. Pushing the H atom toward the dienophile creates an asynchronous TS, which lowers the reaction energy. curs during the distortion of the anthracene reactant and TS are shown in Fig. 4B and fig. S45. We observed that distorting the anthra- cene along OSiC and SiCC angles increases DE. However, bending the top H (CCH) toward the dienophile, from the relaxed geometry of be- tween 180° and 190° to 200°, decreases the DE of Alk, Acr, and MAcr by 2.6, 4.4, and 5.6 kJ·mol−1, respectively, which is consis- tent with the experimentally determined DDG‡ and follows the identical trend MAcr > Acr > Alk > Mal. Only in the case of Mal did bend- ing of the CCH angle increase the DE, most likely because of repulsive electrostatic inter- actions between the bulky dienophile and the SiO2 cluster, although this is still consistent with the smallest DDG‡ observed experimen- tally for Mal. Subsequently, we explored the effect of bending of other angles or bending multiple angles simultaneously (fig. S46 and table S14), and these calculations did not re- sult in substantially different values of DE. This particular bending motion is productive because it increases asynchronicity—the dif- ference in length (DrC–C) between newly formed C–C bonds—of the TS (Fig. 4C). This deforma- tion from the relaxed, nearly synchronous TS of the Acr (DrAcr C(cid:2)C ¼ 0:07 Å) to the distorted, asynchronous TS (DrAcr(cid:2)200 ¼ 0:18 Å) reduces C(cid:2)C the destabilizing activation strain and unfav- orable Pauli repulsion between the reactants, which results in a lower DE (53, 54). Fur- thermore, the predicted degree of the asyn- chronicity of the distorted TSs (table S13) is consistent with the increasing acceleration of the reaction, MAcr > Acr > Alk > Mal. Mechanical force, which acts as uniaxial compression, modifies the potential energy surface of the Diels-Alder reaction differently than does hydrostatic pressure. Under strictly solvothermal conditions (Fig. 5A), DE is equal to the strictly solvothermal activation energy Ea. Hydrostatic pressure (Fig. 5B) acts along the reaction coordinate by reducing the ener- TS), and DDE = gy of the transition state (DEH TS = pDV‡. DV‡ of Diels-Alder reactions DEH under hydrostatic pressure are modest (−20 to −45 cm3·mol−1) (46) because hydrostatic pressure squeezes equally along three axes and does not cause molecular distortion. As such, a p of ~150 MPa is needed for a DDE of −4.4 kJ·mol−1, which is an arbitrarily chosen value of DDE that is in the range of those DDG‡ measured here experimentally. Uniaxial com- pression (Fig. 5C) distorts the reactants, rais- ing their energy by DEd R, and also distorts the transition state to raise its energy by DEd TS. So, under mechanochemical conditions, in which uniaxial compression can cause distortions to R = pDV ‡. Because TS − DEd occur, DDE = DEd the magnitudes for the DV‡ we observed for mechanically distorted Diels-Alder reactions are large (>104 cm3·mol−1), the same DDE of −4.4 kJ·mol−1 can be achieved at p < 1 MPa. We also found that similarities exist (fig. S47) in how force modifies the potential ener- gy surfaces of mechanochemical bond forma- tion (pushing, which causes uniaxial stress) and mechanochemical bond rupture reactions (pulling, which involves uniaxial stress in the opposite direction). In 2007, Hickenboth et al. studied bond rupture in another pericyclic reaction, the ring opening of a benzocyclobutene mechanophore, while using sonication to drive the ring-opening reaction (20). Since then, other pericyclic reactions, including the retro Diels-Alder reaction driven by pulling, have also been examined (17, 22). The Diels-Alder reaction (bond forming) and benzocyclobutene ring-opening–retro Diels-Alder reactions (bond rupture) are all pericyclic reactions, meaning that their regioselectivities and stereoselectiv- ities are governed by the Woodward-Hoffmann rules. The studies by Hickenboth et al. first revealed that a distorted state also precedes the TS and bond rupture in ring opening, and the result of this distorted state is a decrease in DE, where the relationship DDE = DEd R, the same relationship that we observed for mechanochemical bond formation, also applies. Thus, the key to understanding re- gioselectivity and stereoselectivity in mecha- nochemical pericyclic reactions—and bringing mechanochemically driven pericyclic reactions within the Woodward-Hoffmann manifold— necessitates understanding the distortion that uniaxial mechanical stress induces in the re- actants and the TSs. There is, however, an im- portant difference between how pushing and pulling modify the reaction potential energy surface: a Diels-Alder reaction driven by com- pression will follow a distinct trajectory along the potential energy surface than the retro Diels-Alder reaction of its products. As such, the geometries and energies of the TSs and the distorted states of the Diels-Alder reaction and the retro Diels- Alder reaction will be very different. TS − DEd We have used elastomeric tip arrays to con- trol precisely the t and F applied between dienes and dienophiles on a surface to deter- mine k, Ea, DV‡, DG‡, and DDG‡ for the Diels-Alder reaction between immobilized anthracene and four dienophiles that differ in steric and electronic demand. We found that DV‡ is ~1000-fold greater on surfaces under uniaxial compression than under hydrostatic pressure, indicating that the reactions follow different trajectories. The value of tip-based approaches for measuring the kinetic parameters of mech- anochemical bond-breaking reactions is well known, and in this work we show that the techniques are equally valuable for studying bond-forming mechanochemical reactions. These results have important implications for the study, understanding, and wider application and adoption of mechanochemistry. These results also explain that the large DV‡ for mechanochemical reactions occur as a result Zholdassov et al., Science 380, 1053–1058 (2023) 9 June 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. Comparison of reaction trajectories under strictly solvothermal activation, hydrostatic pressure, and uniaxial stress. (A) Reaction potential energy diagram of the Diels-Alder reaction under strictly solvothermal activation. (B) Reaction potential energy diagram of the Diels-Alder reaction under hydrostatic pressure. The arrows indicate the hydrostatic forces that accelerate the reaction. DEH the change in the DE in the hydrostatic mechanism at a p of 150 MPa. (C) Reaction potential energy diagram of the Diels-Alder reaction under uniaxial TS was calculated by using DV‡ of −30 cm3·mol−1 to predict R) and TS deformation energy (DEd mechanochemical activation. The arrows indicate the deformation of the CCH angle on the diene from the solvothermal TS value of 187° to the arbitrarily chosen distorted TS value of 200° and associated reactant deformation energy (DEd TS). All calculations have been performed at M06-2X/6-311+G(d,p) level of theory. The equations for calculating the reaction energy (DE) in each of the three reaction conditions—solvothermal activation, hydrostatic pressure, and mechanical distortion—are provided below the potential energy diagrams. of the proximity to surfaces and uniaxial stress that cause molecular distortion. This distor- tion lowers the reaction energy and drives the reaction through a distinct reaction trajectory, which explains why different isomers are often obtained under F than in solvent (9, 55). This phenomenon—the distortion of molecules near surfaces that occurs under F to lower DG‡— will be relatively independent of molecular structure and suggests that mechanochemical reactivity is more widespread than previously anticipated. As a consequence, these results should encourage the adoption of mechano- chemical methods for sustainable chemical synthesis and for accelerating reactions that are otherwise impractically slow. RE FE RENCES AND N OT ES 1. S. G. Koenig et al., ACS Sustain. Chem.& Eng. 7, 16937–16951 (2019). 2. R. K. Henderson et al., Green Chem. 13, 854–862 (2011). 3. D. R. Joshi, N. Adhikari, J. Pharm. Res. Int. 28, 1–18 (2019). 4. US Energy Information Administration, “2018 Manufacturing Energy Consumption Survey” (US Energy Information Administration, 2021). 5. G.-W. Wang, Chem. Soc. Rev. 42, 7668–7700 (2013). 6. H. P. DeGroot, T. P. Hanusa, Organometallics 40, 3516–3525 (2021). J.-L. Do, T. Friščić, ACS Cent. Sci. 3, 13–19 (2017). 7. 8. T. Friščić, C. Mottillo, H. M. Titi, Angew. Chem. Int. Ed. 59, 1018–1029 (2020). J. G. Hernández, C. Bolm, J. Org. Chem. 82, 4007–4019 (2017). 9. 10. J. M. Andersen, J. Mack, Chem. Sci. 8, 5447–5453 (2017). 11. F. J. Gotor, M. Achimovicova, C. Real, P. Balaz, Powder Technol. 233, 1–7 (2013). 12. D. Crawford et al., Chem. Sci. 6, 1645–1649 (2015). 13. P. A. May, J. S. Moore, Chem. Soc. Rev. 42, 7497–7506 (2013). 14. J. Li, C. Nagamani, J. S. Moore, Acc. Chem. Res. 48, 2181–2190 (2015). 15. A. L. Black, J. M. Lenhardt, S. L. Craig, J. Mater. Chem. 21, 1655–1663 (2011). 16. J. Li et al., J. Am. Chem. Soc. 136, 15925–15928 (2014). 17. J. Wang et al., Nat. Chem. 7, 323–327 (2015). 18. G. R. Gossweiler et al., ACS Macro Lett. 3, 216–219 (2014). 19. K. L. Berkowski, S. L. Potisek, C. R. Hickenboth, J. S. Moore, Macromolecules 38, 8975–8978 (2005). 20. C. R. Hickenboth et al., Nature 446, 423–427 (2007). 21. M. Horst et al., J. Am. Chem. Soc. 143, 12328–12334 (2021). 22. A. R. Sulkanen et al., J. Am. Chem. Soc. 141, 4080–4085 (2019). 23. J. Sung, M. J. Robb, S. R. White, J. S. Moore, N. R. Sottos, J. Am. Chem. Soc. 140, 5000–5003 (2018). 24. J.-M. Seo, I.-Y. Jeon, J.-B. Baek, Chem. Sci. 4, 4273–4277 (2013). 25. L. Gonnet et al., ACS Sustain. Chem.& Eng. 9, 4453–4462 (2021). 26. K. S. McKissic, J. T. Caruso, R. G. Blair, J. Mack, Green Chem. 16, 1628–1632 (2014). 27. I. Fleming, Frontier Orbitals and Organic Chemical Reactions (Wiley, 1976). 28. A. W. Tricker, G. Samaras, K. L. Hebisch, M. J. Realff, C. Sievers, Chem. Eng. J. 382, 122954 (2020). 29. H. Kulla et al., Chem. Commun. 53, 1664–1667 (2017). 30. I. Fleming, Pericyclic Reactions (Oxford Univ. Press, 1998). 31. S. Bian et al., J. Am. Chem. Soc. 135, 9240–9243 (2013). 32. X. Han, S. Bian, Y. Liang, K. N. Houk, A. B. Braunschweig, J. Am. Chem. Soc. 136, 10553–10556 (2014). 33. Y. Hong et al., J. Phys. Chem. C 123, 19715–19724 (2019). 34. C. Huang, H. E. Katz, J. E. West, Langmuir 23, 13223–13231 (2007). 35. X. Liao, A. B. Braunschweig, Z. Zheng, C. A. Mirkin, Small 6, 1082–1086 (2010). 36. X. Liao, A. B. Braunschweig, C. A. Mirkin, Nano Lett. 10, 1335–1340 (2010). 37. L. Huang et al., Small 6, 1077–1081 (2010). 38. J. J. Dudkowski, E. I. Becker, J. Org. Chem. 17, 201–206 (1952). 39. W. R. Vaughan, K. S. Andersen, J. Org. Chem. 21, 673–683 (1956). 40. B. Chen, R. Hoffmann, R. Cammi, Angew. Chem. Int. Ed. 56, 11126–11142 (2017). 41. A. Martini, S. H. Kim, Tribol. Lett. 69, 150 (2021). 42. F. Huo et al., Science 321, 1658–1660 (2008). 43. D. J. Eichelsdoerfer et al., Nat. Protoc. 8, 2548–2560 (2013). 44. A. Loyter, V. Citovsky, R. Blumenthal, Methods Biochem. Anal. 33, 129–164 (1988). 45. C. Walling, H. J. Schugar, J. Am. Chem. Soc. 85, 607–612 (1963). 46. T. Asano, W. J. Le Noble, Chem. Rev. 78, 407–489 (1978). 47. M. G. Evans, M. Polanyi, Trans. Faraday Soc. 31, 875–894 (1935). 48. H. Eyring, J. Chem. Phys. 3, 107–115 (1935). 49. H. Spikes, Friction 6, 1–31 (2018). 50. L. Fang, S. Korres, W. A. Lamberti, M. N. Webster, R. W. Carpick, Faraday Discuss. 241, 394–412 (2023). 51. L. J. Andrews, R. M. Keefer, J. Am. Chem. Soc. 77, 6284–6289 (1955). 52. K. E. Wise, R. A. Wheeler, J. Phys. Chem. A 103, 8279–8287 (1999). 53. F. M. Bickelhaupt, K. N. Houk, Angew. Chem. Int. Ed. 56, 10070–10086 (2017). 54. P. Vermeeren, T. A. Hamlin, F. M. Bickelhaupt, Phys. Chem. Chem. Phys. 23, 20095–20106 (2021). 55. S. Biswas et al., Faraday Discuss. 241, 266–277 (2023). 56. Y. Zholdassov et al., Acceleration of Diels-Alder reactions by mechanical distortion. Dryad (2023); https://doi.org/10.5061/ dryad.0cfxpnw72. 57. Y. Zhao, D. G. Truhlar, Theor. Chem. Acc. 119, 525–525 (2008). 58. M. Linder, T. Brinck, Phys. Chem. Chem. Phys. 15, 5108–5114 (2013). AC KNOWLED GME NTS We thank P. Nautiyal for assistance with acquiring AFM images. We also thank T.-D. Li for assistance with acquiring ToF-SIMS data. Funding: A.B.B., R.W.C., and A.M. are grateful to the National Science Foundation Center for the Mechanical Control of Chemistry (CCI CHE-2023644) for generous support. A.B.B. also acknowledges award DBI-2032176 from the National Science Foundation. This work was carried out in part at the Singh Center for Nanotechnology, which is supported by the NSF National Nanotechnology Coordinated Infrastructure Program under grant NNCI-2025608. Author contributions: Y.S.Z., A.M., R.W.C., M.M., and A.B.B. conceived the research. Y.S.Z., L.Y., S.R.G., A.B., R.W.K., and D.J.V. carried out the experiments. All authors contributed to the writing of the manuscript. Competing interests: The authors declare no competing financial interests. Data and materials availability: All data are available in the supplementary materials. Tables S1 to S14 have been uploaded in more reusable tabular format to the Dryad repository (56). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf5273 Materials and Methods Figs. S1 to S47 Tables S1 to S14 References (59–78) Submitted 27 October 2022; resubmitted 14 February 2023 Accepted 13 April 2023 10.1126/science.adf5273 Zholdassov et al., Science 380, 1053–1058 (2023) 9 June 2023 6 of 6
10.1126_science.adf4930
RES EARCH GEOPHYSICS Ultralow frictional healing explains recurring slow slip events Srisharan Shreedharan1,2*, Demian Saffer1,3*, Laura M. Wallace1,4, Charles Williams4 Plate motion on shallow subduction megathrusts is accommodated by a spectrum of tectonic slip modes. However, the frictional properties and conditions that sustain these diverse slip behaviors remain enigmatic. Frictional healing is one such property, which describes the degree of fault restrengthening between earthquakes. We show that the frictional healing rate of materials entrained along the megathrust at the northern Hikurangi margin, which hosts well-characterized recurring shallow slow slip events (SSEs), is nearly zero (<0.0001 per decade). These low healing rates provide a mechanism for the low stress drops (<50 kilopascals) and short recurrence times (1 to 2 years) characteristic of shallow SSEs at Hikurangi and other subduction margins. We suggest that near-zero frictional healing rates, associated with weak phyllosilicates that are common in subduction zones, may promote frequent, small-stress-drop, slow ruptures near the trench. O ver the past 2.5 decades, a rapidly grow- ing body of observations has revealed a continuum of slip behavior on tectonic plate boundary faults globally, spanning time scales from seconds to years (1). These modes of slip include fast elastodynamic earthquakes, tectonic tremors, low-frequency earthquakes (LFEs), very-low-frequency earth- quakes (VLFEs), slow slip events (SSEs), and aseismic creep. The recognition of these diverse slip modes has sparked a rethinking of the underlying mechanics of earthquakes and fault slip and raised fundamental questions about in situ conditions, rheology, and processes along the subduction interface that govern this spectrum of behavior (1, 2). SSEs, in particular, have emerged as a poten- tially important mode of slip on the shallow subduction zone megathrust. During SSEs, the fault slips slowly over a period of days, weeks, or even months (1), rather than in seconds as in typical elastodynamic earthquakes. In some cases, SSEs may be accompanied by tremors, LFEs, and VLFEs, which radiate seismic energy (1, 3) and are detected seismologically. Numerous hypotheses have been proposed to explain this spectrum of fault slip modes, including highly elevated pore-fluid pressures (4, 5), dilatant strengthening (6), transitional frictional properties (7–10), low-rigidity fault rocks (11), and lithological and geometric hetero- geneities within the fault zone (12). These hy- potheses center around material properties that control the nucleation of an instability on a fault and its ability (or inability) to grow into a dynamic rupture. A relatively unexplored yet 1Institute for Geophysics, Jackson School of Geosciences, University of Texas at Austin, Austin, TX, USA. 2Department of Geosciences, Utah State University, Logan, UT, USA. 3Department of Geological Sciences, Jackson School of Geosciences, University of Texas at Austin, Austin, TX, USA. 4GNS Science, Lower Hutt, New Zealand. *Corresponding author. Email: srisharan.shreedharan@usu.edu (S.S.); demian@ig.utexas.edu (D.S.) equally fundamental control on earthquake occurrence is frictional healing, which describes the ability of a fault to restrengthen and store elastic strain energy between events (2, 10). In the absence of a mechanism for restrengthen- ing, a fault will fail only by aseismic creep. Frictional healing of an earthquake source region is particularly important because it plays a primary role in controlling rupture properties, including stress drop, recurrence interval, and potentially slip mode and amount (2, 10). Measuring the frictional properties of rocks in the fault zones hosting these complex slip behaviors has proved challenging because the SSE source depths are often inaccessible to direct sampling or drilling (13). One notable exception is the Hikurangi subduction zone offshore New Zealand, where regularly recur- ring shallow (<15 km) SSEs are well documented. We report on laboratory measurements of fric- tional healing conducted on samples of rock types involved in slow slip, obtained by drill- ing at the offshore Hikurangi margin. We then integrate these measurements with models of interevent tectonic loading and observa- tions of SSE stress drop. We show that the very low healing rates of fault zone materials along the megathrust, which are typical of clay-rich faults in both subduction zones and other geological settings, are a key ingredi- ent for the occurrence of recurring, frequent, small-stress-drop, slow slip transients. Tectonic setting of the northern Hikurangi margin Westward subduction of the Pacific plate beneath the Australian plate occurs along the northern Hikurangi Trough at 50 to 60 mm/year (14) (Fig. 1, A and B). The sub- ducting plate is primarily composed of the Hikurangi Plateau, a Cretaceous large igneous province, and is overlain by ~1 km of volcani- clastic, pelagic, and siliciclastic sediments (12). The northern Hikurangi margin represents an ideal natural laboratory to investigate the origin of slow earthquake phenomena, owing to the shallow source depth of regularly recurring SSEs, which has enabled a host of high-resolution near-source geophysical investigations, as well as acquisition of host rock samples by drilling (15). The shallow SSEs here originate at depths of 2 to 15 km (12, 13), in a structurally complex region of seamount subduction (Fig. 1C) and intervening areas of high-amplitude seismic reflectivity and low seismic velocity (5, 16), with the latter usually inferred to represent elevated pore- fluid pressure (16). These SSEs propagate to within a few kilometers of, and possibly to, the trench (13); recur every 1 to 2 years; and last for a few weeks (Fig. 1, D and E), during which the plate interface experiences slip rates of ~1 cm/day (15, 17). The SSEs are character- ized by small stress drops (<10 kPa) and a range of energy release equivalent to that for moment magnitude (Mw)~6 to 7 earthquakes (18). Microseismicity and tremor have been documented along and above the plate inter- face during and after the shallow SSEs (15). The megathrust in this region has also hosted two tsunami earthquakes (long-duration elasto- dynamic earthquakes with slow rupture veloc- ities of ~1 km/s) in 1947 (19). Thus, the shallow megathrust here is home to a rich variety of slip modes that appear to interact in complex ways in space and time. International Ocean Discovery Program (IODP) Expedition 375 sampled Cretaceous volcani- clastic conglomerates overlying the subduct- ing Hikurangi Plateau at Site U1520, located ~10 km east of the trench (Fig. 1, A to C). Seismic imaging indicates that these materials are entrained along and eventually form the shal- low megathrust here (12, 20). The margin is characterized by low heat flow, with tempera- tures on the megathrust expected to remain <150°C at a distance of ≳70 to 100 km from the trench (21). Thus, the frictional and rheologi- cal properties of the volcaniclastic material recovered at Site U1520 are likely represen- tative of those along the shallow megathrust (to ~10- to 12-km depth), well into the source region of regularly recurring SSEs. Frictional healing, fault failure, recurrence, and stress drop Frictional healing describes the ability of a material to restrengthen frictionally over time and is a prerequisite for repeating stick-slip (seismic) failure of faults because it allows interseismic coupling and elastic strain accu- mulation (22). Healing is thought to reflect a time-dependent increase in asperity contact area at the micro- to nanoscopic scales (22). In the laboratory, frictional healing is commonly measured via slide-hold-slide (SHS) tests (23) and typically increases log-linearly with time (22). Healing (b) is usually reported as the Shreedharan et al., Science 379, 712–717 (2023) 17 February 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E A B D 40 30 20 10 ) s y a d ( n o i t a r u d p i l S 4 0 S G 7 0 S G 6 0 S G 0 1 S G 1 1 S G E 4 1 S G 3 1 S G 80 60 40 20 ) a P k ( p o r d s s e r t S 2004 2008 2012 Year C Fig. 1. Tectonic setting of the Gisborne area shallow slow slip events. (A) Bathymetric map of the northern Hikurangi margin showing the trench (bold black curve), depth to the plate interface (dashed contours), slip contours of the 2014 SSE in millimeters (black, as labeled), location of the 05CM-04 seismic line (black line), and locations of IODP expedition 375 drill sites (red circles) [modified from (12)]. Hypocenter of the 1947 tsunami earthquake (blue star), tremor activity (small yellow circles), and microseismicity (white stars) are also shown. (B) Regional tectonic map showing the Pacific plate subducting beneath the Australian plate, forming the northern Hikurangi margin. (C) Interpretation of the 05CM-04 seismic line [modified from (12)] showing the plate interface (bold black line), active upper-plate splay faults (red lines), and location and depth extent of IODP drill site U1520. The blue star at a depth of ~6 km represents the hypocenter of the 1947 tsunami earthquake. (D) Slip duration and (E) Spatiotemporally averaged stress drop of the quasiperiodic SSEs rupturing the Gisborne patch from 2002 to 2016 (table S2). rate of increase in fault strength (or friction) per logarithmic (10-fold) increase in time. We analyze the SHS experiments conducted by (10) on Cretaceous volcaniclastic sediments from Site U1520 (Fig. 2A). Modal compositions of these samples (table S1) (24) determined from x-ray diffraction indicate that they contain >55 wt % clay minerals and that the clay is dominantly smectite. These experiments were performed on water-saturated gouges under an effective normal stress of 25 MPa, pore pres- sures of 5 to 10 MPa, and for driving velocities of 1 to 30 mm/s. These experimental boundary conditions were selected to best represent appropriate conditions of stress and sliding velocity at shallow SSE source depths along the subduction megathrust. The healing rates inferred for these materials (b < 10−4, Fig. 2B) represent some of the lowest values of fault healing reported for fault rocks and are more than 10 to 100 times below those com- monly reported for carbonates and tectosili- cates (25). Our observations are consistent with near-zero healing rates documented in phyllosilicate-rich synthetic and natural fault gouges, including smectite and serpentinite (25, 26) (Fig. 2B). We compare these frictional healing rates to average stress drops reported for repeating shallow SSEs that ruptured the Gisborne patch (18) (table S2) (Figs. 1 and 2). Seismic evidence indicates that the materials participating in shallow SSEs here may be compositionally similar to the input materials that were fric- tionally tested (12, 20). Although we have incomplete knowledge of the exact in situ effective stress (seff) along the shallow SSE source fault (2, 12, 15), indirect observations strongly suggest significant overpressure along the plate interface (2, 5). To honor these obser- vations and also allow for reasonable uncer- tainty in seff, we convert the friction drop from SHS experiments to a stress drop (Fig. 2C) by considering a range of pore-fluid overpressure values along the plate interface of l = 0.8 to 0.95 (where l = Pf/Pl, and Pf and Pl are the fluid and lithostatic pressures, respectively) (23) (fig. S1). The laboratory data for the frictionally weak volcaniclastic conglomerates are in excellent agreement with the geodetically defined stress drops for the Gisborne shallow SSEs (Fig. 2C). Together, these analyses suggest near-zero fric- tional restrengthening rates. In comparison, repeating earthquakes along the Calaveras fault exhibit substantially higher stress drops (1 to 10 MPa) that are strongly correlated with repeat- ing earthquake recurrence intervals (Fig. 2C), increasing by nearly 2 MPa per 10-fold in- crease in time (27). SHS experiments on the predominantly quartzofelspathic material recovered from the Calaveras fault indicate correspondingly high frictional healing rates of ~0.007 (28), at least two orders of mag- nitude higher than the healing rates we re- port for the Hikurangi megathrust materials (Fig. 2B). Along the San Andreas Fault, ex- perimentally estimated frictional healing rates on materials recovered from drilling are also in agreement with both the recur- rence and stress drop of small repeating earth- quakes along bounding faults and with the creeping behavior of the central deforming zone (25, 26, 28). Competition between loading and healing on the shallow subduction megathrust To further explore the role of the experimen- tally observed near-zero healing rates in govern- ing event recurrence and strain accumulation and release, we consider the balance between tectonic loading and healing rates. We com- pute loading rates along the plate interface using a two-dimensional (2D) numerical model of the margin that incorporates heter- ogeneous elastic properties, which are well constrained by geophysical surveys (29) (Fig. 3, A and B) (23). We then compare these loading rates to the event-averaged stress drop and recurrence intervals of the ~10-year sequence of observed Gisborne slip events (Fig. 3C). The latter assumes that all of the strain energy accumulated during an interevent period is released during slip. The numerical model predicts tectonic stress- ing rates ranging from 30 kPa/year at 95 km from the trench (near the downdip edge of the SSE region) to 0.3 kPa/year at 30 km (the updip end of the likely SSE source region), with an average loading rate of ~4.5 kPa/year over the outer ~65 km of the megathrust (Fig. 3B). Event- based estimates of stress accumulation rates using observed stress drops and recurrence intervals for each SSE range between 1 to 30 kPa/year, with an average of ~9 kPa/year across the seven events (Fig. 3, B to C). The two estimates of tectonic loading rates are in excel- lent agreement and suggest that on average, the loading rate along the shallow megathrust is ~4 to 10 kPa/year. The slip distributions for most offshore Gisborne SSEs are poorly con- strained, with the exception of the most recent Shreedharan et al., Science 379, 712–717 (2023) 17 February 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E slide hold A reslide μss μpeak p5391 U1520 28R5 μss μpeak μss μss μ 2 0 . 0 µ i , t n e c fi f e o c n o i t c i r F p5392 U1520 38R5 200 Loadpoint displacement (μm) 300 100 400 Laboratory experiments SSEs and repeaters λ = 0.80 λ = 0.90 λ = 0.95 7) 2 ( ult fa s era v ala C e d a c e a/d P M 1-2 Gisborne SSEs 102 104 Hold time or recurrence interval (s) 106 108 C 8 6 4 2 0 ) a P M ( Δ , p o r d s s e r t S 100 0 10 B 8 6 4 2 0 -2 100 3 - 0 1 x µ Δ , g n i l a e h l a n o i t c i r F Δμ peak , p5391 Δμ peak , p5392 SAFOD CDZ and cuttings (26, 28) Δμss (10) (25) fault 7x10-3 Calaveras = β β=10-3 0-2 1 = β β = 10-4 101 102 Hold time (s) 103 104 Fig. 2. Frictional healing constrained from laboratory experiments. (A) Overlay of multiple SHS protocols with different durations (3 to 3000 s) for two phyllosilicate-rich sediment samples from Site U1520. mss represents the pre- and posthold steady-state values of the friction coefficient, and mpeak represents the location of the (near-zero) peak friction upon reshear after a hold. (B) The change in friction (healing) between mpeak and mss shows little or no correlation with hold time. Error bars quantify electrical noise in the load cell and represent the limit of resolution of the friction drops. Fiducial lines show three different healing rates representing a range of values reported in prior studies (25), as well as values encompassed by the experiments reported here. Gray line represents the friction drops from SHS experiments conducted on samples from the Calaveras fault (25), where seismic repeaters have been observed. Pink area shows the range of values reported from SHS experiments conducted on smectite-rich San Andreas Fault Observatory at Depth (SAFOD) Central Deforming Zone (CDZ) cuttings (26, 28). Light blue region represents changes in steady-state friction coefficient from pre- to posthold reported by (10). (C) Changes in stress drop versus recurrence interval for the Gisborne SSEs and versus hold time for the experimental data. Stress drops for experimental data are calculated as the product of friction drop and effective stress for three different degrees of overpressure. Gray triangles represent stress drops estimated from the seismic source for Calaveras repeaters (27). events, for which seafloor geodetic data were available (13). This makes a precise compari- son of the stress drops in the events with the predicted loading rates for an identical fault region difficult. Nonetheless, our results are consistent with similarly small loading rates in the outer forearc at the Nankai margin (30). Frictional restrengthening increases logarith- mically with time (as b × seff), whereas the shear stress borne by the fault increases, to first order, linearly with time as a function of tectonic loading, which in turn is a function of rock and fault elastic properties, geometry, and plate convergence rate (Figs. 2 and 3). At a given location, and in the absence of addi- tional loading—for example, associated with slip of adjacent fault patches—the fault reaches a threshold for failure when the accumulated stress exceeds the fault strength (or when the interevent loading exceeds the amount of fric- tional restrengthening; Fig. 4, A to C). In this simplified framework, the time to failure repre- sents the recurrence interval and the stress at failure is the average stress drop of a quasiperi- odic event. Here, we combine our numerically determined loading rates (Fig. 3) with a re- strengthening model to understand the size and recurrence of time- and slip-predictable events both through time at specific locations (Fig. 4, A to C) and as a function of position along the megathrust (Fig. 4, D to F). Both tectonic loading and fault healing increase with depth along the plate inter- face (Figs. 2 and 3). As noted above, without strong constraints on seff, we consider a range of likely pore-fluid pressure ratios based on geological and geophysical inferences (2, 5) (Fig. 4 and fig. S2). We evaluate the competi- tion between the loading rate and healing— and thus the expected recurrence and stress drops of repeating fault failure events—as a function of distance from the trench by con- sidering the spatial distribution of excess strength at different times. We do so for a range of likely pore-fluid pressures (and by proxy seff), as well as for a range of b that are consistent with our data. Because our exper- imental data provide only an upper bound on b, we explore values in the range of b = 3 × 10−5 to 3 × 10−6, representing up to 1.5 orders of magnitude lower than this upper bound. We illustrate the competition between load- ing and restrengthening at three representa- tive locations along the plate interface, i.e., 25, 50, and 75 km from the trench (Fig. 4, A to C) for an effective stress profile corresponding to l = 0.9. At 75 km from the trench, well within the SSE source region, the plate interface experiences tectonic loading of ~3.5 kPa/year (Fig. 3B). For our upper bound of b (3 × 10−5), we predict a recurrence interval of 2.3 years, with events having average stress drops of ~8 kPa; the lowest value of b (3 × 10−6) that we consider yields events with recurrence in- tervals of 3 to 4 months and <1 kPa stress drops (Fig. 4A). Closer to the trench, at a distance of 25 km, the loading rate on the plate interface is much lower (0.3 kPa/year). As a result, for our upper bound on b, we predict considerably longer times to failure (15 years) and stress drops of 4 kPa (Fig. 4C). Our lower bound on b yields recurrence intervals of ~1.3 years and stress drops of ~0.4 kPa. Low b (<10−4), coupled with the slow loading rates on the shallow plate interface, suggest that the fault is likely to fail in repeating, fre- quent (1- to 3-year) events, with small maxi- mum total slip (on the order of a few to ~10 cm), on the basis of the slip deficit accrued during the interevent time and on low stress drops (on the order of a few to ~10 kPa). Although b and seff are not perfectly constrained, other combina- tions of these parameters (fig. S2) yield similar outcomes. The observed recurrence and stress drops in the Gisborne SSEs are most consistent with scenarios presenting modest-to-high fluid overpressure (l > 0.8), although near-lithostatic pressures are not required (fig. S2). Also, Be- cause the rate of fault restrengthening (b × seff) scales with effective normal stress, this frame- work provides a mechanism, in addition to conditional frictional stability (2, 7), that links Shreedharan et al., Science 379, 712–717 (2023) 17 February 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E ) m k ( t h p e D 0 2 4 6 8 10 12 ) r y / a P k ( e t a r i g n d a o L 102 101 100 10-1 102 101 100 10-1 Slow slip rupture region Subducted seamount 1947 tsunami earthquake Gisborne events GS04 GS11 GS10 GS14 GS07 GS13 GS06 SSE source region 90 80 70 60 50 30 Distance from trench (km) 40 A B C Numericalmodel Model average in SSE source 20 10 0 Fig. 3. Inter-SSE loading rates in the northern Hikurangi margin. (A) An illustration of the 05CM-04 line showing upper-plate splay faults (red curves), the hypocenter of the 1947 tsunami earthquake (red star), a subducting seamount, and the likely shallow SSE rupture region (thick yellow bars). (B) Annual tectonic loading rates (for convergence at 50 mm/year) along the 05CM-04 seismic transect, computed from 2D finite-element numerical model in Pylith (29). Loading rates increase from ~0.15 kPa/year near the trench to ~30 kPa/year at a depth of 15 km. (C) Tectonic loading rates constrained from the Gisborne SSEs as the ratio of average stress drops and interevent times. The black (~4.5 kPa/year) dashed line shows average loading rates from models. elevated pore pressure with frequent small- stress-drop events, as is commonly postulated or inferred for SSEs (5). We evaluate the spatial variation of this competition in more detail by considering the strength excess along the plate interface, defined as the difference between strength (or amount of restrengthening) and accumu- lated tectonic stress at different times (Fig. 4, D to F), i.e., after 1, 2, and 3 years, for an effective stress profile corresponding to l = 0.9 (fig. S1). A positive strength excess represents a fault that has healed more than it has been loaded tectonically, indicating that it is still far from failure. Conversely, a negative strength excess indicates a fault that has already failed because it has experienced more tectonic stress- ing than restrengthening. For the case where b < 3 × 10−5, the strength excess is nearly zero to a depth of ~8.5 km (corresponding to distances ~0 to 50 km from the trench), indicating that the ultralow healing rates of the SSE source at such shallow depths may allow it to fail fre- quently (i.e., on a time scale of months or shorter) and with very small (sub-kilopascal) stress drops (Fig. 4, D and E). Further landward from the trench and deeper, the strength excess reaches zero after 1 to 3 years, which is in good agreement with the recurrence intervals of the shallow SSEs here. An additional complexity in extrapolating healing rates to natural faults arises because some studies indicate that healing (b) may vary with effective stress, temperature, and lithology (10, 31, 32). This variation is cap- tured across the suite of cases that we ex- plore for values of b ranging from 3 × 10−6 to 3 × 10−5 (Fig. 4, A to F). Some studies in- dicate that experimental healing rates could increase with larger hold times (>3000 s) via pressure solution or other mechanisms, par- ticularly in strong quartzofeldspathic minerals (32, 33), whereas others indicate that healing rates may reduce over long time scales, par- ticularly for clay-rich materials as the quality and quantity of contact area becomes saturated (28). Here, in accordance with (25, 28), we consider a constant healing rate in mapping from the experimental to natural SSE time scales. In the constant b model (Fig. 4G, left), stress drops are relatively constant at ~6 to 10 kPa regardless of the downdip location of the SSE source, but there is considerable variability in predicted recurrence intervals. If b increases with depth (Fig. 4G, right), we predict frequent (<9- to 14-months recur- rence) SSEs with stress drops of <1 kPa near the trench, and larger, less-frequent SSEs further downdip (~10 kPa, 1- to 3-year re- currence) at 10- to 15-km depth. The latter are detectable by traditional onshore geo- detic instruments (15), whereas the former may be detectable only by offshore borehole observatories near the trench (34). These small, frequent, slow transients, if real, raise the possibility of a spectrum of slip spanning from quasiperiodic creep or undulating slip on the shallowest reaches of the megathrust. These transients originate near the trench because, although loading is small, healing is negligible. This would result in a continuum from recurring larger SSEs farther from the trench transitioning to more frequent, smaller events near the trench. Moreover, depending on the timescale and nature of observations, these small, slow, and frequently repeating transients could present themselves as aseismic creep, particularly in the absence of continuous, near-field, and high-precision offshore moni- toring; in the limit as healing approaches zero, slip will occur via stable creep. In other words, when healing and loading are equivalently small, such as near the trench, aseismic creep and SSEs may be indistinguishable from one another. The small stress drops and small total slip predicted by low healing rates are also phys- ically consistent with fault failure via slow earthquakes or SSEs as opposed to rapid slip (35). Small stress drops and slip limit the stored elastic energy available for the accel- eration of slip, therefore favoring slip in slow motion and over a long duration (35, 36). Additionally, SSEs are thought to involve substantial ductile deformation that extends beyond a narrow slip zone (37), which would dissipate a substantial amount of energy off- fault, further reducing the energy available for propagating a dynamic rupture on-fault (38), and in turn, keeping any instability slow. Taken together, the recurring small stress drop events of <10 kPa (18) and velocity-neutral frictional stability of the volcaniclastic materials (10) indicate an inability of the SSE source region to heal between SSEs and an inability to nucleate large, dynamic earthquakes char- acterized by large fault slip and rapid slip velocities. Shreedharan et al., Science 379, 712–717 (2023) 17 February 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E t RI 10 A 8 6 4 2 x = 75 km B x = 50 km C x = 25 km g β = 3 x 10-5 Δ β = 10-5 β = 3 x 10-6 g g g g ) a P k ( g n i l a e h r o g n d a o i l i c n o t c e T l a n o i t c i r f 0 1 2 3 4 5 0 ) a P k ( s s e c x e h t g n e r t S 20 10 0 -10 -20 β = 3 x 10-6 D Gisborne SSE source Not failed E a 1yr 2 3 1yr 2 3 Failed 2 4 Time (years) 6 β = 10-5 b Not failed Failed 0 4 8 12 16 β = 3 x 10-5 F 1yr 2 3 Not failed c Failed a P k 0 1 s s e r t s r a e h S Distance landward from trench (km) G Constant healing with depth, β = 3x10-5 Increasing healing with depth x = 75 km x = 75 km, β = 3x10-5 x = 50 km x = 50 km, β = 10-5 x = 25 km x = 25 km, β = 3x10-6 Time 5 years Fig. 4. A framework to quantify competition between tectonic loading and frictional healing. Tectonic loading as shown in Fig. 3 (black lines) and fault restrengthening (colored curves) as a function of time, for three different assumed healing rates (3 × 10−6, purple; 1 × 10−5, orange; and 3 × 10−5, green) at (A) 75 km, (B) 50 km, and (C) 25 km from the trench. Strength excess as a function of distance from the trench, after 1 year (dotted curve), 2 years (dashed curve), and 3 years (solid curve) for healing rates of (D) 3 × 10−6, (E) 1 × 10−5, and (F) 3 × 10−5 [same color scheme as in panels (A) to (C)]. (G) Simplified models of recurrence and stress drop for two scenarios of depth-varying healing, showing shear-stress variations over time. Constant healing with depth (left, light to dark blue) results in small-stress-drop events with large recurrence intervals near the trench; increasing healing rate with depth (right, gray to black) results in very-small-stress-drop events (~1 kPa) every few weeks or months. Effective stress corresponds to l = 0.90 for all panels. The prediction that a near-zero strength excess is attained simultaneously over a broad region of the plate interface (e.g., Fig. 4, D and E) also implies that the shallow megathrust reaches failure over a broad region. This is consistent with observations of shallow SSEs having a large slip patch with small total slip (35), because the low loading rates combined with low healing rates never allow for a large accumulation of slip deficit. In contrast, or- dinary earthquakes with similar amounts of slip (a few centimeters) commonly nucleate over much smaller areas (35). In addition to hosting SSEs, the northern Hikurangi margin (12, 13) and other shallow plate boundaries (2, 34, 39, 40) also host tsunami earthquakes, microseismicity, LFEs, VLFEs, and tremor. Carbonates, which are characterized by substantially larger b (10) (compare to Fig. 2) and capable of nucleating instabilities even at modestly elevated temper- atures (41, 42), are also present in the North Hikurangi SSE source region, though with less abundance than the volcaniclastics (10, 12, 41). This heterogeneity of lithology may provide one explanation for the diverse slip behaviors observed along the margin. For example, it raises the possibility that infrequent tsunami earthquakes could nucleate locally where strong and unstable materials such as carbon- ates are embedded and/or abundant within the broader SSE region (12). In this scenario, such ruptures could then propagate through the weak, phyllosilicate-rich volcanoclastic-rich portions of the megathrust (43), producing tsunami earthquakes (19). Alternatively, small carbonate- or quartz-filled fractures along the décollement may be loaded during an SSE. This increased loading, coupled with the ele- vated healing rates of these materials, could result in tremorgenic failure that is spatiotem- porally colocated with SSE slip (34, 44). We integrate experimental measurements of healing and healing rates on relevant mega- thrust materials, a numerical model of fault loading, and observed shallow SSE source properties. Taken together, they indicate that low healing rates on weak faults are a key ingredient in the generation of frequent, low- stress-drop, recurring slip events with small amounts of total slip, such as those observed at the northern Hikurangi margin and many plate boundaries globally. Because the low rates of healing limit the stored strain and stress (and strain energy) available in the system, materials with low healing rates are likely to host SSEs, creep, and other slow earth- quake phenomena. Furthermore, because re- strengthening is controlled by the product of frictional healing and effective normal stress, our results highlight an additional path- way by which high pore pressure (or low ef- fective stress) may directly influence the size, frequency, and nature of failure, and ultimate- ly modulate the mode of failure in a host of tectonic and geologic settings, particularly at shallow depths (1). Weak phyllosilicate min- erals, which typically exhibit low healing rates, are common along shallow subduction mega- thrusts (2), raising the possibility that shallow SSEs and near-trench creep may be the norm at subduction zones and that the compara- tively limited observations of shallow SSEs at subduction zones globally may simply be a consequence of our inability to detect them with land-based geodetic networks. REFERENCES AND NOTES 1. Z. Peng, J. Gomberg, Nat. Geosci. 3, 599–607 (2010). 2. D. M. Saffer, L. M. Wallace, Nat. Geosci. 8, 594–600 (2015). 3. G. C. Beroza, S. Ide, Annu. Rev. Earth Planet. Sci. 39, 271–296 (2011). 4. D. Wei, Y. Sandwell, Y. Fialko, J. Geophys. Res. 114, B07402 (2009). 5. D. Bassett, R. Sutherland, S. Henrys, Earth Planet. Sci. Lett. 389, 1–13 (2014). 6. P. Segall, M. Rubin, M. Bradley, R. Rice, J. Geophys. Res. 115, 7. B12305 (2010). J. R. Leeman, D. M. Saffer, M. M. Scuderi, C. Marone, Nat. Commun. 7, 11104 (2016). 8. H. S. Rabinowitz et al., Geochem. Geophys. Geosyst. 19, 2973–2990 (2018). 9. K. Im, D. Saffer, C. Marone, J.-P. Avouac, Nat. Geosci. 13, 705–710 (2020). 10. S. Shreedharan et al., Geochem. Geophys. Geosyst. 23, e2021GC010107 (2022). 11. Y. Ito, K. Obara, Geophys. Res. Lett. 33, 9 (2006). 12. P. M. Barnes et al., Sci. Adv. 6, eaay3314 (2020). 13. L. M. Wallace et al., Science 352, 701–704 (2016). 14. L. M. Wallace, J. Beavan, R. McCaffrey, D. Darby, J. Geophys. Res. 109, B12406 (2004). 15. L. M. Wallace, Annu. Rev. Earth Planet. Sci. 48, 175–203 (2020). 16. R. Bell et al., Geophys. J. Int. 180, 34–48 (2010). 17. L. M. Wallace, J. Beavan, J. Geophys. Res. 115, B12402 (2010). 18. M. J. Ikari et al., Geochem. Geophys. Geosyst. 21, e2019GC008717 (2020). Shreedharan et al., Science 379, 712–717 (2023) 17 February 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 19. R. Bell, C. Holden, W. Power, X. Wang, G. Downes, . Earth 32. H. Yasuhara, C. Marone, D. Elsworth, J. Geophys. Res. 110, Planet. Sci. Lett. 397, 1–9 (2014). B06310 (2005). 20. A. C. Gase, N. L. Bangs, H. J. A. Van Avendonk, D. Bassett, 33. M. J. Ikari, B. M. Carpenter, C. Vogt, A. J. Kopf, Earth Planet. S. A. Henrys, Geology 50, 1145–1149 (2022). 21. A. Antriasian et al., Geophys. J. Int. 216, 1177–1190 (2019). 22. J. H. Dieterich, B. D. Kilgore, Pure Appl. Geophys. 143, 283–302 (1994). 23. Materials and methods are available as supplementary materials. 24. M. B. Underwood, “Data report: Clay mineral assemblages within biocalcareous and volcaniclastic inputs to the Hikurangi subduction zone, IODP Expedition 372B/375 Sites U1520 and U1526, offshore New Zealand.” Proceedings of the International Ocean Discovery Program 372B/375 (2021). 25. B. M. Carpenter, M. J. Ikari, C. Marone, J. Geophys. Res. Solid Earth 121, 1183–1201 (2016). 26. M. M. Scuderi, B. M. Carpenter, Geophys. J. Int. 231, 290–305 (2022). 27. C. Marone, J. E. Vidale, W. L. Ellsworth, Geophys. Res. Lett. 22, 3095–3098 (1995). 28. B. M. Carpenter, D. M. Saffer, C. Marone, Geology 40, 759–762 (2012). 29. C. A. Williams, L. M. Wallace, Geophys. Res. Lett. 45, 7462–7470 (2018). 30. K. Tamaribuchi, M. Ogiso, A. Noda, J. Geophys. Res. Solid Earth 127, e2022JB024403 (2022). 31. G. C. McLaskey, A. M. Thomas, S. D. Glaser, R. M. Nadeau, Nature 491, 101–104 (2012). Sci. Lett. 450, 164–172 (2016). 34. E. Araki et al., Science 356, 1157–1160 (2017). 35. E. E. Brodsky, J. Mori, Geophys. Res. Lett. 34, 16 (2007). 36. Y. Ida, Bull. Seismol. Soc. Am. 63, 959–968 (1973). 37. Å. Fagereng et al., Geology 47, 872–876 (2019). 38. A. Chounet, M. Vallée, M. Causse, F. Courboulex, Tectonophysics 733, 148–158 (2018). 39. M. Radiguet et al., J. Geophys. Res. 117, B040305 (2012). 40. Y. J. Tan, D. Marsan, Sci. Adv. 6, eabb2489 (2020). 41. A. M. Eijsink, M. J. Ikari, Geochem. Geophys. Geosyst. 23, e2022GC010369 (2022). 42. R. M. Kurzawski, M. Stipp, A. R. Niemeijer, C. J. Spiers, J. H. Behrmann, Nat. Geosci. 9, 717–722 (2016). 43. H. S. Rabinowitz, H. M. Savage, P. J. Polissar, C. D. Rowe, J. D. Kirkpatrick, Nat. Commun. 11, 533 (2020). 44. Å. Fagereng, F. Remitti, R. H. Sibson, Tectonophysics 510, 381–386 (2011). 45. S. Shreedharan, Data for Hikurangi SSE and Healing Paper, Zenodo (2022); https://doi.org/10.5281/ zenodo.7246267. ACKN OWLED GMEN TS D.M.S. acknowledges funding from a US Science Support Program post-expedition award and the Scott Petty Jr. Director’s Chair endowment. L.W. acknowledges funding from the New Zealand MBIE Endeavor Research Fund (contract C05X1605) and the Marsden Fund (grant 20-GNS-001). Author contributions: Conceptualization: S.S., D.M.S. Methodology: S.S., D.M.S., L.W., and C.W. Investigation: S.S., D.M.S., L.W., and C.W. Visualization: S.S., D.M.S. Funding acquisition: S.S., D.M.S., and L.W. Writing – original draft: S.S. Writing – review and editing: S.S., D.M.S., L.W., and C.W. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data are available in the main text, supplementary materials, or from Zenodo (45). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf4930 Materials and Methods Supplementary Text Figs. S1 to S3 Tables S1 and S2 References (46–53) We thank L. Lavier and D. C. Bolton for insightful discussions. Funding: S.S. was funded by the UTIG Palisades Postdoctoral Fellowship. Submitted 25 October 2022; accepted 25 January 2023 10.1126/science.adf4930 Shreedharan et al., Science 379, 712–717 (2023) 17 February 2023 6 of 6
10.1126_science.adf5341
Climate-driven changes in soil conditions alter hibernation We used long-term climate records of air and soil temperature from two field sites, Toolik (68°38′ N, 149°38′ W; elevation 719 m) and Atigun (68°27′ N, 149°21′ W; elevation 812 m), in the Alaskan Arctic to document recent envi- ronmental change. The dataset durations were as follows: air temperatures at Toolik were measured from fall 1993 to spring 2020, Toolik soil temperatures from fall 1993 to spring 2019, and Atigun soil temperatures from fall 2002 to spring 2019. These measures were paired with hibernation records collected using biologgers (Toolik, fall 1996 to spring 2019; Atigun, fall 1999 to spring 2021) (15) to evaluate the physiological impact of recent climate change on arctic ground squirrels (see the supplementary materials). The use of biologgers that measured abdominal and/or skin temperature allowed us to record detailed information about physiological and phenological events during hibernation (Fig. 1, fig. S1, and table S1) from 199 free-living arctic ground squirrel individuals over 25 years. Average annual air temperatures increased from 1994 to 2020 (b = 0.070, t = 2.552, P = 0.02; Fig. 2A). This was driven by increases in winter temperatures, because seasonal analyses revealed increases in winter temperatures (b = 0.119, t = 2.375, P = 0.025; fig. S2), but no annual trends for spring, summer, or fall (see the supplementary materials and fig. S2). Dates of soil freeze, measured at 1-m depth adjacent to hibernacula, shifted to be ~4 days/decade later in the fall (n = 446, b = 0.409, t = 2.851, df = 383, P = 0.0045; Fig. 2B), and minimum soil temperatures in winter increased by al- most 2°C/decade (n = 366, b = 0.185, t = 6.922, P < 0.001; Fig. 2C). Further, dates of soil thaw in summer advanced by ~0.3 days/decade RES EARCH HIBERNATION Climate change is altering the physiology and phenology of an arctic hibernator Helen E. Chmura1,2*, Cassandra Duncan3, Grace Burrell3, Brian M. Barnes1, C. Loren Buck4, Cory T. Williams5* Climate warming is rapid in the Arctic, yet impacts to biological systems are unclear because few long-term studies linking biophysiological processes with environmental conditions exist for this data-poor region. In our study spanning 25 years in the Alaskan Arctic, we demonstrate that climate change is affecting the timing of freeze-thaw cycles in the active layer of permafrost soils and altering the physiology of arctic ground squirrels (Urocitellus parryii). Soil freeze has been delayed and, in response, arctic ground squirrels have delayed when they up-regulate heat production during torpor to prevent freezing. Further, the termination of hibernation in spring has advanced 4 days per decade in females but not males. Continued warming and phenological shifts will alter hibernation energetics, change the seasonal availability of this important prey species, and potentially disrupt intraspecific interactions. C limate change is particularly rapid in the Arctic (1), where systematic warming is reducing sea ice extent, altering hydro- logical cycles, thawing permafrost, increas- ing shrubs, and changing the timing of key seasonal events (phenological shifts) (2). Despite the rapid pace of climate change in the Arctic, it is a relatively data-poor region (3), and few long-term records combining physical records of climate change and physiological responses of organisms exist [but see (4, 5)]. Further, although changes in the spring and summer have received considerable attention, recent work has called for more research into the consequences of warmer and wetter win- ters (6). Winter conditions shape life histories, because many animal species have evolved strat- egies such as seasonal migration or dormancy to cope with prolonged periods of low food availability (7). In resident species that hiber- nate, climate change could lead to changes in energy expenditure and overwinter sur- vival. Energy requirements increase markedly when hibernacula temperatures drop below an animal’s thermal set point [near freezing for ground squirrels (8)], because animals must produce heat to prevent tissue damage and death (9). Winter temperatures may also contribute to the regulation of spring life his- tory events (10, 11), and phenological shifts can have important ecological repercussions if they result in mismatches such that histor- ically synchronous interactions within or among species are no longer temporally aligned (12, 13). 1Institute of Arctic Biology, University of Alaska Fairbanks, Fairbanks, AK 99775, USA. 2Rocky Mountain Research Station, United States Forest Service, Missoula, MT 59801, USA. 3Department of Biology and Wildlife, University of Alaska Fairbanks, Fairbanks, AK 99775, USA. 4Department of Biological Sciences, Northern Arizona University, Flagstaff, AZ 86011, USA. 5Department of Biology, Colorado State University, Fort Collins, CO 80523, USA. *Corresponding author. Email: cory.williams@colostate.edu (C.T.W.); helen.chmura@usda.gov (H.E.C.) As part of a 25-year field study of arctic ground squirrels (Urocitellus parryii) in the Alaskan Arctic, we evaluated the impacts of climate change on this high-latitude mammalian hi- bernator, focusing on changes in hibernation physiology and emergence phenology. Unlike hibernators in most temperate or montane re- gions, which sequester themselves in hibernacula that remain above freezing, arctic ground squir- rels overwintering in frozen soils must defend themselves against large thermal gradients (or differences between ambient and body tem- peratures) while torpid using nonshivering thermogenesis (thermogenic torpor) (9, 14) (Fig. 1). We demonstrate that significant warm- ing in ambient air and soil (hibernacula) tem- peratures is altering hibernation phenology and the duration of thermogenic torpor in this arctic species. Fig. 1. Seasonal changes in arctic ground squirrel body temperature and soil temperature of the hiber- naculum. Abdominal body temperature of a hibernating arctic ground squirrel (blue line) and temperature of the surrounding hibernaculum (soil temperature at 1-m depth adjacent to the burrow entrance, dashed brown line) from a site near Toolik Field Station. Squirrels expend energy to maintain their body temperature above that of the hibernaculum (“thermogenic torpor”) such that their brain temperatures never drop below 0°C (9). Chmura et al., Science 380, 846–849 (2023) 26 May 2023 1 of 3 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Long-term changes in temperature and hibernation physiology. Over the past 25 years at Toolik Field Station in the Alaskan Arctic, mean annual air temperatures increased (A), soil freeze date occurred later (sampling at 1-m depth, adjacent to hibernacula) (B), and minimum overwinter soil temperatures increased (C). Concurrently, the onset of thermogenic torpor occurred later in the fall [(D); shown for females only]. Although the date at which female ground squirrels initiated hibernation in the fall did not change over time (E), the date that they ended hibernation in the spring became earlier (F). Lines represent mean results of linear mixed effects models, and shaded regions represent 95% confidence intervals of the mean. Points represent raw data, and colors represent study sites (Atigun or Toolik). For all animal data, data points are plotted against the year in which an individual initiated hibernation. (n = 262, b = –0.36, 95% credible interval, –0.57 to –0.15). In combination, this resulted in an ~10-day reduction in the annual duration that soil was frozen at 1-m depth over the study period. In arctic ground squirrels, the date on which core body temperatures first decreased below 0°C each hibernation season, i.e., the date when squirrels first became thermogenic during torpor (9), was delayed by ~15 days per decade over the 25-year course of our study (n = 256, b = 1.484, t = 4.800, df = 91, P < 0.001; Fig. 2D), with no difference seen between sexes (b = 3.039, t = 1.041, df = 161, P = 0.299). However, other phenological events did not show the same pattern. Neither adult males (n = 122, b = –0.354, t = –1.454, df = 32, P = 0.156) nor adult females (n = 182, b = –0.163, t = –0.819, df = 71, P = 0.416; Fig. 2E) changed the date at which hibernation began in fall. Adult females, how- ever, ended hibernation earlier in spring (n = 166, b = –0.389, t = –3.592, df = 60, P < 0.001; Fig. 2F) at a rate of ~4 days/decade. As females ended hibernation earlier over the 25-year period, ani- mal mass measured during the first 2 weeks after emergence increased (n = 91, b = 3.472, 95% credible interval 0.196 to 6.781; see the sup- plementary materials and fig. S3). Males did not shift spring phenology (n = 120, b = –0.192, t = –1.103, df = 30, P = 0.279). No directional changes were observed in parturition timing (n = 91, b = –0.003, t = –1.343, df = 28, P = 0.190). Discussion Winter plays a fundamental role in determin- ing species’ range limits and local popula- tion dynamics (6, 16). We show that warming in the AlaskanArctic has altered seasonal freeze- thaw dynamics of the soil active layer and thus the thermoregulatory patterns of arctic ground squirrels during hibernation. Delayed freeze of the soil active layer near hibernacula is delaying the onset of thermogenic torpor and presumably reducing energy expenditure and overwinter weight loss. Further, we found sex differences in phenological responses to climate change, with females advancing their spring active season by 10 days over 25 years and males showing no change. Warmer winter conditions and the short- ening of the hibernation season in females have the potential to affect the survival prob- ability of free-living arctic ground squirrels. There are several general mechanisms through which this could occur, including changes in energetics and predation exposure. First, de- laying thermogenic torpor will decrease the total amount of time that arctic ground squir- rels spend defending body temperature dur- ing hibernation, which will reduce overwinter energy expenditures (9, 17) and increase spring mass in females, as we have demonstrated in this study. Additionally, changes in winter tem- peratures will alter the energetic costs of torpor bouts and episodic interbout arousals (8) that characterize hibernation in small mammals. Arctic ground squirrels spend most of the hibernation season generating heat to avoid freezing, unlike most hibernators in more mod- erate climates, which can safely thermoconform because the surrounding soil temperatures remain above freezing during torpor (18). Re- duced intensity and/or duration of thermo- genesis caused by warmer conditions would Chmura et al., Science 380, 846–849 (2023) 26 May 2023 2 of 3 RES EARCH | R E S E A R C H A R T I C L E allow arctic ground squirrels to conserve en- ergy and potentially increase winter survival, particularly among vulnerable, energy-limited age classes such as juveniles (19, 20). Alter- natively, warmer conditions associated with climate change could shorten the hibernation season and increase the number of days that arctic ground squirrels are active above ground, increasing mortality rates because of increased exposure to predators (21). Thus, the conse- quences of climate change for hibernators will likely be heterogeneous (22). Although our study focused on a system in which warmer air temperatures lead to warmer hibernacula temperatures and likely reduced energy expen- diture, in other systems, climate change may reduce snow depth, which can decrease burrow temperatures and decrease energetic savings through hibernation (20). We found sex differences in phenological flexibility, with female arctic ground squirrels, but not males, terminating hibernation earlier. Other studies suggest that female ground squirrels are less sensitive to the direct effects of temperature (23, 24) and instead are respon- sive to temperature-driven changes in spring snow cover conditions (25). Thus, female flex- ibility appears to allow them to match ener- getic demands with the environment, and we expect that earlier snowmelt will also corre- spond with earlier vegetation green-up (26). As winters continue to warm, sex differences in phenological shifts may lead to disrupted intersexual interactions. For example, dur- ing one extremely warm spring in eastern Canada, female Richardson’s ground squir- rels (Urocitellus richardsonii) became sexually receptive before most males were physiologi- cally prepared to mate, resulting in a pheno- logical mismatch between the sexes (27). In the short term, mismatches such as these could affect population reproductive rates, and over longer time scales, continued warming in the Arctic may be a strong selective force resulting in evolutionary changes in male phenology. The consequences of climate change may be direct, such as the changes in hibernation physiology and phenology that we report here, as well as indirect. If the population size and temporal availability of arctic ground squirrels above ground are altered by climate change, then this could have cascading indirect effects on diverse tundra predators. Understanding the impact of climate change on species such as the arctic ground squirrel will aid assessments of how arctic food webs will function in a rapidly warming world, and research linking physiol- ogy and phenology to demographic responses is an important component of understanding community responses to climate change. RE FERENCES AND NOTES F. Pithan, T. Mauritsen, Nat. Geosci. 7, 181–184 (2014). 1. J. E. Box et al., Environ. Res. Lett. 14, 045010 (2019). 2. J. E. Hobbie et al., Ambio 46 (Suppl. 1), 160–173 (2017). 3. 4. E. Post, M. C. Forchhammer, Philos. Trans. R. Soc. Lond. B Biol. Sci. 363, 2369–2375 (2008). 5. G. Gauthier et al., Philos. Trans. R. Soc. Lond. B Biol. Sci. 368, 20120482 (2013). 6. C. M. Williams, H. A. L. Henry, B. J. Sinclair, Biol. Rev. Camb. Philos. Soc. 90, 214–235 (2015). 7. K. Wilsterman, M. A. Ballinger, C. M. Williams, Funct. Ecol. 35, 11–31 (2021). 8. M. M. Richter et al., Physiol. Biochem. Zool. 88, 81–89 (2015). 9. B. M. Barnes, Science 244, 1593–1595 (1989). 10. S. P. Caro, S. V. Schaper, R. A. Hut, G. F. Ball, M. E. Visser, PLOS Biol. 11, e1001517 (2013). 11. H. E. Chmura, C. T. Williams, Horm. Behav. 144, 105215 (2022). 12. S. J. Thackeray et al., Glob. Chang. Biol. 19, 3568–3580 (2013). 13. M. D. Burgess et al., Nat. Ecol. Evol. 2, 970–975 (2018). 14. C. L. Buck, B. M. Barnes, J. Mammal. 80, 1264–1276 (1999). 15. C. T. Williams, B. M. Barnes, C. L. Buck, Comp. Biochem. Physiol. A Mol. Integr. Physiol. 202, 53–62 (2016). 16. D. Gudex-Cross et al., Glob. Ecol. Biogeogr. 31, 1366–1380 (2022). 17. S. A. Karpovich, Ø. Tøien, C. L. Buck, B. M. Barnes, J. Comp. Physiol. B 179, 691–700 (2009). 18. T. Ruf, F. Geiser, Biol. Rev. Camb. Philos. Soc. 90, 891–926 (2015). 19. C. Lenihan, D. Van Vuren, Can. J. Zool. 74, 297–302 (1996). 20. C. Rézouki et al., J. Anim. Ecol. 85, 761–773 (2016). 21. C. Turbill, S. Prior, Funct. Ecol. 30, 1366–1372 (2016). 22. C. P. Wells, R. Barbier, S. Nelson, R. Kanaziz, L. M. Aubry, Ecography 2022, e06056 (2022). 23. B. M. Barnes, D. Ritter, in Life in the Cold: Ecological, Physiological, and Molecular Mechanisms (CRC Press, 1993), pp. 119–130. 24. H. E. Chmura et al., Integr. Comp. Biol. 62, 1012–1021 (2022). 25. C. T. Williams et al., Am. Nat. 190, 854–859 (2017). 26. I. P. La Puma, T. E. Philippi, S. F. Oberbauer, Remote Sens. Environ. 109, 225–236 (2007). 27. C. E. Kucheravy et al., Sci. Rep. 11, 21684 (2021). 28. Data for: H. E. Chmura, C. Duncan, G. Burrell, B. M. Barnes, C. L. Buck, C. T. Williams, Climate change is altering the physiology and phenology of an arctic hibernator, Dryad (2023); https://doi.org/10.5061/dryad.pzgmsbcqq. AC KNOWLED GME NTS We thank J. T. Moore, F. Kohl, M. Richter, and many additional students and technicians who worked countless hours in the field to support this work. We are also grateful for 30 years of outstanding support from staff and facilities at the Toolik Field Station, which made this research possible. Funding: This work was supported by the National Science Foundation (grant IOS-1558056 to C.T.W. and C.L.B.); the National Science Foundation (grant IOS-1558160 to B.M.B.); a University of Alaska Centennial Postdoctoral Fellowship (H.E.C.); the National Institute of General Medical Sciences of the National Institutes of Health [Institutional Development Award (IDeA) grant P20GM103395 to B.M.B. and C.D.]; and a University of Alaska Fairbanks URSA undergraduate research award (G.B.). Author contributions: Conceptualization: H.E.C., C.T.W.; Data curation: H.E.C., C.D., G.B.; Formal analysis: H.E.C., C.D., G.B.; Funding acquisition: B.M.B., C.L.B., C.T.W.; Investigation: H.E.C., C.D., G.B., C.L.B., C.T.W.; Methodology: H.E.C., B.M.B., C.L.B., C.T.W.; Project administration: C.T.W.; Resources: B.M.B., C.L.B., C.T.W.; Supervision: C.T.W.; Visualization: H.E.C.; Writing – original draft: H.E.C., C.T.W.; Writing – review and editing: H.E.C., B.M.B., C.L.B., C.T.W. Competing interests: The authors declare no competing interests. Data and materials availability: The datasets and code generated during the study are available in the Dryad repository (28). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf5341 Materials and Methods Supplementary Text Figs. S1 to S3 Table S1 References (29–39) MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. Submitted 31 October 2022; accepted 25 April 2023 10.1126/science.adf5341 Chmura et al., Science 380, 846–849 (2023) 26 May 2023 3 of 3
10.1126_science.adi1847
RES EARCH POPULATION DYNAMICS Boom-bust cycles in gray whales associated with dynamic and changing Arctic conditions Joshua D. Stewart1*, Trevor W. Joyce2,3, John W. Durban3,4, John Calambokidis5, Deborah Fauquier6, Holly Fearnbach4, Jacqueline M. Grebmeier7, Morgan Lynn3, Manfredi Manizza8, Wayne L. Perryman3, M. Tim Tinker9,10, David W. Weller3 Climate change is affecting a wide range of global systems, with polar ecosystems experiencing the most rapid change. Although climate impacts affect lower-trophic-level and short-lived species most directly, it is less clear how long-lived and mobile species will respond to rapid polar warming because they may have the short-term ability to accommodate ecological disruptions while adapting to new conditions. We found that the population dynamics of an iconic and highly mobile polar-associated species are tightly coupled to Arctic prey availability and access to feeding areas. When low prey biomass coincided with high ice cover, gray whales experienced major mortality events, each reducing the population by 15 to 25%. This suggests that even mobile, long-lived species are sensitive to dynamic and changing conditions as the Arctic warms. T he Bering and Chukchi seas in the Pacific Arctic are extremely productive shallow basins (1–3) that support seasonal for- aging opportunities for a wide variety of migratory and Arctic-associated taxa (4). The Pacific Arctic food web is characterized by ice-associated algal growth during spring and early summer, which is transported to the benthos through decay and sinking of particu- late organic carbon (3). This tight pelagic- benthic coupling historically resulted in some of the most productive nearshore benthic sys- tems in the world (3), attracting migrants from throughout the Pacific and supporting large populations of marine species (4, 5). As the Arctic has rapidly warmed, sea ice retreat has occurred progressively earlier in the spring, and the Bering and Chukchi seas have remained ice free for longer in the autumn (6). This has resulted in increased water-column productivity (7, 8) but has reduced the amount of particulate organic carbon that reaches the sea floor through pelagic-benthic coupling that is dependent on sinking ice-associated algae (5). In addition, decreased sea ice cover allows stronger current-driven flow over the shallow basins of the Pacific Arctic, reducing the quan- tity of finer-sediment grain size within the 1Ocean Ecology Lab, Marine Mammal Institute, Department of Fisheries, Wildlife and Conservation Sciences, Oregon State University, Newport, OR, USA. 2Ocean Associates, Arlington, VA, USA. 3Marine Mammal and Turtle Division, National Oceanic and Atmospheric Administration (NOAA) Southwest Fisheries Science Center, La Jolla, CA, USA. 4Sealife Response, Rehabilitation and Research (SR3), Des Moines, WA, USA. 5Cascadia Research Collective, Olympia, WA, USA. 6Office of Protected Resources, National Marine Fisheries Service, Silver Spring, MD, USA. 7Chesapeake Biological Laboratory, University of Maryland Center for Environmental Science, Solomons, MD, USA. 8Geosciences Research Division, Scripps Institution of Oceanography, University of California San Diego, La Jolla, CA, USA. 9Nhydra Consulting, Halifax, NS, Canada. 10Ecology and Evolutionary Biology, University of California, Santa Cruz, Santa Cruz, CA, USA. *Corresponding author. Email: joshua.stewart@oregonstate.edu benthos that support habitat for tube-building amphipods, which have some of the highest lipid content of benthic crustaceans (9, 10). Collectively, these impacts have driven changes to the structure of Arctic benthic communities, which may translate into impacts on higher- trophic-level species that migrate seasonally to access these foraging hotspots (5, 9, 10). Eastern North Pacific gray whales (Eschrichtius robustus) undertake one of the longest mam- malian migrations between wintering areas in Baja California, Mexico, and summer feeding areas in the Bering and Chukchi seas to take advantage of these highly concentrated ben- thic prey resources (11). Gray whales have spe- cialized baleen plates adapted to suction feeding in soft sediments and are the only baleen whale to feed primarily on benthic prey (11). Al- though they are capable of feeding on pelagic zooplankton, the diet of gray whales feeding in the Arctic is dominated by benthic crustaceans— in particular, ampeliscid amphipods—that are found in abundance in shallow Arctic basins (12). Estimates of pre-whaling population sizes range from 15,000 to 30,000 individuals for the eastern North Pacific gray whale popula- tion, based on population models fitted to esti- mates from abundance surveys combined with commercial and aboriginal harvest data (13). Genetic estimates of prehistoric abundance are much higher, ranging from ~75,000 to 120,000 individuals (14), although this likely included the now endangered western North Pacific population and may reflect a larger carrying capacity supported by increased ben- thic habitat availability during the Last Glacial Minimum (15). Commercial whaling in the lagoons of Baja California and throughout the North Pacific depleted the eastern North Pacific gray whale population to fewer than 5000 individuals by the early 1900s (13). A rapid and sustained post-whaling increase in abundance led to the delisting of the popula- tion from the Endangered Species Act in 1994 and is widely viewed as an iconic example of successful conservation and species recovery (16). The status and stability of eastern North Pacific gray whales has come into question as the population experienced two major docu- mented mortality events in 1999–2000 and 2019–2022 (17, 18). In response to the first mortality event in 1999, there was speculation that the population may have reached its carrying capacity and was suffering from density-dependent effects on survival (19). In light of fluctuations in reproductive output and a second major mortality event two decades later, many studies have proposed that variable and changing Arctic conditions may be drivers of eastern North Pacific gray whale population dynamics (12, 20–22). Arctic sea ice extent has been proposed as a contributor to gray whale vital rates—especially reproduction—by physically restricting access to summer feeding areas (20, 22, 23). However, in recent years previously identified relation- ships between gray whale reproduction and Arctic sea ice extent have begun to decouple (22, 23), and variability in sea ice has been insufficient to explain mortality rates (20). Eastern North Pacific gray whales have the most complete long-term abundance and demo- graphic data available for any large whale species, and we leveraged these extensive data- sets to examine environmental drivers of pop- ulation dynamics not possible in other species. We combined time series of gray whale abun- dance, reproduction, nutritive condition, and strandings spanning more than half a century into a population dynamics model to esti- mate annual carrying capacity for the pop- ulation. We show that this annual carrying capacity is well explained by ice-mediated access to the population’s primary foraging grounds in the Arctic and biomass of benthic crustaceans. The observed boom-bust cycles in gray whale abundance and vital rates suggest that as large whales recover from post-whaling depletion, their populations may become in- creasingly governed by environmental con- straints and climate variability. Results and discussion We combined 31 estimates of eastern North Pacific gray whale abundance over 54 years (1968 to 2022) (24), 30 estimates of calf pro- duction over 42 years (1980 to 2022) (22, 25), 1391 records of stranded gray whales on the United States coastline over 48 years (1974 to 2022), and 1334 body condition measure- ments over 32 years (1987 to 2019) (26) into an integrated population dynamics model that estimates annual abundance, birth rates, and mortality rates. The model uses evidence of human interactions in stranded gray whales to estimate proportional hazards of anthropo- genic and natural contributions to mortality. Stewart et al., Science 382, 207–211 (2023) 13 October 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Population dynamics of eastern North Pacific gray whales. (A) Gray whales have experienced major fluctuations in abundance after an initial post- whaling recovery, including three major declines beginning in 1987, 1999, and 2019. (B to E) These declines and subsequent recoveries in the 1990s and 2000s were associated with synchronous changes in (B) births and (C) mortality, as well as changes in nutritive condition in (D) southbound and (E) northbound migrating whales. Black points in (A) and (B) indicate the median estimated abundance and calf pro- duction from visual surveys, with standard errors of model estimates (vertical bars). Black points in (D) and (E) indicate the mean values of body condition measurements from each survey year and the standard deviation of observations (vertical bars). In (A) to (E), the black lines indicate the median of the posterior distribution of model- estimated values, and the shaded regions indicate the 95% posterior credible intervals. In addition, the model estimates both the long- term carrying capacity (K), as well as an annually varying carrying capacity (Kt) that reflects year-to-year variation in the strength of negative density dependence as determined by environ- mental covariates and stochastic effects. We considered three Arctic time series as candi- date covariates for annual gray whale carrying capacity: (i) access to feeding grounds, defined as the number of days with <50% sea ice cover on the historic gray whale foraging grounds in the Chirikov basin and southern Chukchi Sea (1979 to 2021) (23, 27); (ii) benthic infaunal crustacean biomass, averaged over the same foraging hotspots as sea ice access (1971 to 2019) (28); and (iii) zooplankton density es- timated by using a global ocean ecosystem model that includes the entire Arctic Ocean ecosystem, averaged over gray whale foraging hotspots (1992 to 2020) (29). The data and population model are described in detail in the Data sources and Integrated population model sections of the supplementary materials. The eastern North Pacific gray whale pop- ulation has experienced three major mortality events, each resulting in reductions of 15 to 25% of total abundance within the half-century of nearly continuous monitoring, representing extraordinarily high periodic mortality rates for a long-lived vertebrate (Fig. 1). These mortality events were associated with peaks in reported strandings during the 1999–2000 and 2019–2022 periods. The 1987–1989 abundance decline is the largest in magnitude but was not asso- ciated with an increase in strandings, likely because reporting structures and survey effort to detect strandings were expanded and im- proved substantially beginning in 1990. How- ever, this major impact to the population is also reflected in the poorest recorded body condi- tion of the survey history in 1988, falling rapidly from very good condition in 1987 (Fig. 1D). The population dynamics model estimated low an- nual carrying capacities (Kt) of approximately 10,000 individuals during each of these die- offs (Fig. 2A), indicating that Arctic foraging grounds periodically experience major disrup- tions, limiting the number of whales that they can support. These fluctuations in annual carrying capacity were represented in mortality rates, body condition, and most strongly in birth rates, which had the greatest propor- tional change with varying carrying capacity (fig. S5). On the basis of anthropogenic injury rates in stranded whales, model-estimated an- thropogenic mortality rates remained low and stable, whereas natural mortality rates varied substantially and peaked during major die- offs, suggesting direct human impacts such as vessel strikes and entanglements in fishing gear are not the primary drivers of mortality in this population. The maximum birth rate estimated by the model was 0.111 (95% credible intervals 0.108 to 0.114). The realized annual birth rate ranged from a low of 0.0046 in 1998 (0.0024 to 0.0076) to a high of 0.085 in 1975 (0.062 to 0.102). Within the span of calf production ob- servations (1994–2022), the minimum birth rate was 0.007 in 2000 (0.004 to 0.01), and the maximum was 0.082 in 2004 (0.069 to 0.09). The minimum estimated mortality rate was 0.011 (0.009 to 0.014). The realized annual mortality rate ranged from a low of 0.019 in 1975 (0.014 to 0.027) to a high of 0.13 in 1988 (0.099 to 0.162). During the three major mortal- ity events, median estimated mortality rates were 0.13 and 0.079 (in 1988 and 1989); 0.065 and 0.099 (in 1999 and 2000); and 0.092, 0.089, 0.061, and 0.067 (from 2019 to 2022). Model-estimated mean body condition was lowest in 1988 [median 0.162, 95% confidence interval (CI) 0.158 to 0.166], 2000 (0.165, 0.163 to 0.168), and 2020 (0.167, 0.163 to 0.170). The highest estimated body condition was in 1975 (0.184, 0.181 to 0.187), although there were no photogrammetric measurements before 1987. The 3 years with highest estimated body con- dition and corresponding condition measure- ments were 2013 (0.181, 0.180 to 0.183), 2012 (0.181, 0.179 to 0.183), and 1997 (0.181, 0.179 to 0.182). The estimated northbound body Stewart et al., Science 382, 207–211 (2023) 13 October 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Drivers of eastern North Pacific gray whale carrying capacity. (A) Esti- mated annual carrying capacity (Kt) from the population dynamics model, with reference lines at 25,000 (dashed line) and 10,000 (dotted line). (B) Estimated ice access anomaly, which is the Z-scored number of days with 50% or lower ice cover on gray whale feeding grounds. (C) Estimated crustacean biomass anom- aly, which is the Z-scored mean grams of carbon of benthic crustaceans on key gray whale feeding grounds. (D) Decline in benthic crus- tacean per capita biomass from 1970 to 2019, showing the relationship each sampling year between ben- thic crustacean abundance and biomass in grams of carbon (gC). In (A) to (C), the black lines indicate the median of the posterior distribution of estimates, and the shaded regions indicate the 95% posterior credible intervals. condition scaling factor was 0.922 (0.913 to 0.930), indicating an ~8% decline in body con- dition between southbound and northbound measurements. The estimated long-term average K was 22,062 (18,967 to 24,725). This long-term average is lower than the median of annual Kt values (24,500, 95% CI 21,771 to 27,797), which is to be expected given that it is the arithmetic mean outcome of a stochastic process and thus re- flects the effects of environmental variability on expected abundance (30). We found a significant positive relationship between benthic crustacean biomass and carry- ing capacity (99.9% probability slope > 0), no relationship with zooplankton density (39.2% > 0), and a high probability of a positive relation- ship with sea ice access (93.5% > 0). With the zooplankton density covariate eliminated from the model, both crustacean biomass (100% > 0) and sea ice access (96.2% > 0) had significant positive relationships with carrying capacity. This suggests that the ability of the eastern North Pacific gray whale population to physi- cally access key feeding areas, in combination with in situ prey availability, explains fluctua- tions in body condition, reproduction, and mortality. The three major mortality events occurred during periods of simultaneous low crustacean biomass and restricted access to feeding areas (Fig. 2). In 2010, a rapid decrease in crustacean biomass but a period of average ice access led to a depression in birth rates and a modest decrease in abundance but not a major mortality event. The onset of the 2019 mortality event appears to have been driven initially by low crustacean biomass and exacer- bated by a steep reduction in access to feeding areas over the following 2 years. The decision to model gray whale popula- tion dynamics by applying annual covariate effects to carrying capacity (K), rather than the population’s intrinsic growth rate (r), is uncommon. Although in theory either model formulation could be used to explain fluctua- tions in abundance and vital rates, we believe that applying covariate effects to carrying ca- pacity better reflects biological reality. The Bering and Chukchi seas are the primary feed- ing area for virtually all eastern North Pacific gray whales, suggesting that the quality and quantity of prey in these areas will have a greater impact on vital rates when there is high intraspecific competition at higher levels of gray whale abundance. This is supported empiri- cally by our estimates of population growth rate relative to abundance. Mean population growth rates were significantly higher at low than at high abundance levels, and major busts (annual declines of >9 to 10%) only occurred when the gray whale population was at high abundance (fig. S9), which supports the existence of density- dependent controls on vital rates. By applying covariate effects to carrying capacity, we simul- taneously account for environmental conditions and the effects of negative density depen- dence (31). In addition, this avoids a scenario in which, in a model that applies covariate effects to r instead of K, the population exceeds a sta- tionary carrying capacity but continues to grow because of positive covariate effects on growth rate. Instead, our estimated annual carrying capacity (Kt) captures short-term fluctuations in the strength of density dependence and can be interpreted as an abstract parameter corres- ponding to the expected equilibrium abundance if environmental conditions remained fixed at the values recorded during that year (32). Over the past 50 years, the per capita bio- mass of benthic infaunal crustaceans has de- clined precipitously (Fig. 2D and fig. S3), and the three major gray whale mortality events coincided with periods of low per capita bio- mass, which translated into low total crusta- cean biomass. This decline in per capita biomass is most likely associated with species distri- bution shifts of benthic amphipods and other Stewart et al., Science 382, 207–211 (2023) 13 October 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E crustaceans. As ice cover decreases in response to rapid Arctic warming, current speed in the Chirikov basin has increased, leading to larger sediment grain size and reduced particulate organic carbon reaching the seafloor (5). These conditions favor smaller amphipods with lower lipid content over the lipid-rich, tube-building ampeliscid amphipods that historically domi- nated the shallow basins of the Bering and Chukchi seas (10). This regime shift has likely contributed to declining per capita biomass of gray whale prey, which despite steady or in- creasing prey abundance has resulted in lower overall available biomass (fig. S3). The combined effect of sea ice cover and benthic productivity on gray whale population dynamics has driven major boom-bust cycles, including two modern booms in abundance that may have exceeded preexploitation levels (13). High benthic biomass and prey quality in the late 1970s and early 1980s supported al- most 25,000 gray whales, contributing to their delisting from the US Endangered Species Act. More recently, rapid Arctic warming in re- sponse to climate change increased access to feeding areas (Fig. 2B), supporting a sustained increase in gray whale abundance over the past decade (Fig. 1A). Although recent Arctic warming may have provided sufficient benefit to the population to counteract decreasing benthic biomass over the short term, the outlook for benthic prey quality is not favorable. Rising water column and bottom water tem- peratures and projected decoupling of pelagic and benthic productivity caused by retreating sea ice will likely lead to continued declines in Arctic benthic crustacean biomass (5). Access to feeding areas reached a peak of 266 days in 2019, which is presumably approaching a point of diminishing returns given that the spe- cies migrates to Mexico each winter. Poleward shifts in gray whale feeding locations have already been documented, which likely reflect the declining quality and shifting distribution of their preferred prey (12). Future declines in benthic biomass will likely drive decreases in gray whale carrying capacity that cannot be offset by continued increases in ice access. Reports of gray whales shifting their Arctic feeding distribution and targeting pelagic prey (12) suggest that they may have the ability to compensate for these changing conditions to some extent, but our results suggest that any ongoing behavioral adaptations have thus far been insufficient to prevent major mortality events. Eastern North Pacific gray whales are the most closely monitored large whale species, with records of abundance, reproduction, mortality, and condition spanning more than half a cen- tury. The abundance of most large whale species remains far below pre-whaling levels (33, 34), which limits our understanding of the dyna- mics and behavior of whale populations as they approach carrying capacity and become increasingly governed by density-dependent processes. By contrast, gray whales have re- covered rapidly from post-whaling lows to num- bers that may approach or exceed pre-whaling levels and have low rates of direct human mortality, providing a rare window into the possible natural fluctuations of large whale populations. The periodic mortality events and major population swings that we report are surprising for a long-lived vertebrate that must by definition have high average survival rates to facilitate longevity. However, whales achieve their immense body sizes by feeding on large quantities of low-trophic-level prey (35), which may make them sensitive to oceanographic and environmental fluctuations. The feeding- fasting cycles associated with migratory baleen whales may also increase their susceptibility to environmental perturbations. Gray whales migrate more than 15,000 km each year and rely on a 4- to 5-month feeding season to sup- port a majority of their energetic requirements for the year. This strategy may place them at a physiological threshold at which disruptions to their food supply translate into major im- pacts to vital rates—a pattern that may be widespread across migratory whales and may become more pronounced as species and pop- ulations recover to their pre-whaling abun- dances. Climate-driven ocean warming is expected to have profound impacts on ocean circulation, upwelling strength, and primary production (36, 37), which may in turn have major implications for large whale population dynamics and viability through predator-prey interactions (34). RE FERENCES AND NOTES 1. R. C. Highsmith, K. O. Coyle, Nature 344, 862–864 (1990). 2. A. M. Springer, C. P. McRoy, M. Flint, Fish. Oceanogr. 5, 3. 205–223 (1996). J. M. Grebmeier, L. W. Cooper, H. M. Feder, B. I. Sirenko, Prog. Oceanogr. 71, 331–361 (2006). 4. K. J. Kuletz et al., Prog. Oceanogr. 136, 175–200 5. (2015). J. M. Grebmeier et al., Science 311, 1461–1464 (2006). 6. H. P. Huntington et al., Nat. Clim. Chang. 10, 342–348 (2020). 7. K. R. Arrigo, G. L. van Dijken, Prog. Oceanogr. 136, 60–70 (2015). 8. K. M. Lewis, G. L. van Dijken, K. R. Arrigo, Science 369, 9. 198–202 (2020). J. M. Grebmeier, K. E. Frey, L. W. Cooper, M. Kędra, Oceanography 31, 136–151 (2018). 10. J. M. Grebmeier, S. E. Moore, J. E. Overland, K. E. Frey, R. Gradinger, Eos 91, 161–162 (2010). 11. G. C. Pike, J. Fish. Res. Board Can. 19, 815–838 (1962). 12. S. E. Moore et al., PLOS ONE 17, e0265934 (2022). 13. A. E. Punt, C. Allison, G. Fay, J. Cetacean Res. Manag. 6, 121–132 (2004). 14. S. E. Alter, E. Rynes, S. R. Palumbi, Proc. Natl. Acad. Sci. U.S.A. 104, 15162–15167 (2007). 15. N. D. Pyenson, D. R. Lindberg, PLOS ONE 6, e21295 (2011). 16. P. J. Clapham, S. B. Young, R. L. Brownell, Mammal Rev. 29, 37–62 (1999). 17. B. J. Le Boeuf, H. Pérez-Cortés M., J. Urbán R., B. R. Mate, F. Ollervides U., J. Cetacean Res. Manag. 2, 85–99 (2000). 18. F. Christiansen et al., Mar. Ecol. Prog. Ser. 658, 237–252 (2020). 19. S. E. Moore et al., Mar. Mamm. Sci. 17, 954–958 (2001). 20. J. R. Brandon, A. E. Punt, “Assessment of the eastern stock of North Pacific gray whales: incorporating calf production, sea- ice and strandings data,” paper SC/61/AWMP2 presented to the International Whaling Commission (IWC) Scientific Committee (2009). 21. A. E. Punt, P. R. Wade, “Population status of the eastern North Pacific stock of gray whales in 2009,” US Department of Commerce, National Oceanic and Atmospheric Administration (NOAA) Technical Memo NMFS-AFSC-207, 43 (2009). 22. W. L. Perryman, T. Joyce, D. W. Weller, J. W. Durban, Mar. Mamm. Sci. 37, 448–462 (2020). 23. T. W. Joyce et al., Mar. Ecol. Prog. Ser. 709, 141–158 (2023). 24. T. Eguchi, A. R. Lang, D. W. Weller, “Abundance and migratory phenology of Eastern North Pacific gray whales 2021/2022,” US Department of Commerce, NOAA Technical Memorandum NMFS-SWFSC-668 (2022). 25. T. Eguchi, A. R. Lang, D. W. Weller, “Eastern North Pacific gray whale calf production 1994-2022,” US Department of Commerce, NOAA Technical Memorandum NMFS-SWFSC-667 (2022). 26. W. L. Perryman, M. S. Lynn, J. Cetacean Res. Manag. 4, 155–164 (2002). 27. G. Gailey et al., Sci. Rep. 10, 1553 (2020). 28. J. M. Grebmeier, L. W. Cooper, Benthic macroinfaunal and dominant taxa samples collected from Northern Bering Sea to Chukchi Sea, 1970–2019 (2023); https://arcticdata.io/ catalog/view/doi%3A10.18739%2FA24T6F480. 29. D. Carroll et al., J. Adv. Model. Earth Syst. 12, e2019MS001888 (2020). 30. R. C. Lewontin, D. Cohen, Proc. Natl. Acad. Sci. U.S.A. 62, 1056–1060 (1969). 31. J. Roughgarden, Am. Nat. 109, 713–736 (1975). 32. D. G. Heckel, J. Roughgarden, Proc. Natl. Acad. Sci. U.S.A. 77, 7497–7500 (1980). 33. C. Scott Baker, P. J. Clapham, Trends Ecol. Evol. 19, 365–371 (2004). 34. V. J. D. Tulloch, É. E. Plagányi, C. Brown, A. J. Richardson, R. Matear, Glob. Change Biol. 25, 1263–1281 (2019). 35. M. S. Savoca et al., Nature 599, 85–90 (2021). 36. O. Hoegh-Guldberg, J. F. Bruno, Science 328, 1523–1528 (2010). 37. G. C. Hays, A. J. Richardson, C. Robinson, Trends Ecol. Evol. 20, 337–344 (2005). 38. J. D. Stewart et al., stewart6/ENPGW-IPM: Data and Code for Stewart et al. Boom-bust cycles in gray whales. Zenodo (2023); https://doi.org/10.5281/zenodo.8201214. AC KNOWLED GME NTS We thank past and present members of the Working Group for Marine Mammal Unusual Mortality Events; the Gray Whale Unusual Mortality Event Investigative Teams; as well as the US, Canadian, and Mexico marine mammal stranding network responders. We thank the NOAA Office of Marine and Aviation Operations and D. LeRoi for their support of drone flights to measure whale body condition. We thank J. Baker for feedback on analyses and interpretation. The scientific results and conclusions and any views or opinions expressed herein are those of the authors and do not necessarily reflect the views or policies of the US government, its agencies, or any of the included organizations. Funding: J.D.S. was supported by a National Academies NRC Research Associateship and the Oregon State University Marine Mammal Institute Research Endowment. Abundance, calf production, and aerial photogrammetry surveys were supported by NOAA, US Department of Commerce. Drone photogrammetry surveys at Piedras Blancas, California, from 2015 to 2019 were supported by SeaLife Response, Rehabilitation, and Research (SR3) and NOAA, with facility and property use provided by the Bureau of Land Management, US Department of the Interior. J.M.G. was supported for benthic time series sampling for crustaceans and associated fauna through multiple awards, most recently the US National Science Foundation Office of Polar Programs Stewart et al., Science 382, 207–211 (2023) 13 October 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E (OPP 1917469) and the National Oceanic and Atmospheric Administration Arctic Research Program (CINAR 25984.02). Author contributions: Conceptualization: J.D.S. and J.W.D. Funding acquisition: W.L.P., J.W.D., D.W.W., and H.F. Investigation: All authors. Methodology: J.D.S., J.W.D., W.L.P., D.W.W., and M.T.T.; Data curation: T.W.J., J.W.D., J.C., D.F., H.F., J.M.G., M.L., M.M., W.L.P., and D.W.W. Formal analysis: J.D.S. and M.T.T. Visualization: J.D.S. Writing – original draft: J.D.S. Writing – review and editing: All authors. Competing interests: The authors declare no competing interests. Data and materials availability: All data and code required to reproduce the analyses presented in the main text and online supplementary materials are available online through Zenodo (38). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi1847 Materials and Methods Figs. S1 to S9 References (39–55) MDAR Reproducibility Checklist Submitted 8 April 2023; accepted 16 August 2023 10.1126/science.adi1847 Stewart et al., Science 382, 207–211 (2023) 13 October 2023 5 of 5
10.1126_science.ade3332
RES EARCH R E S E A R C H A R T I C L E ◥ RESPIRATORY COMPLEXES Structural basis of mammalian respiratory complex I inhibition by medicinal biguanides Hannah R. Bridges1*, James N. Blaza1,2, Zhan Yin1, Injae Chung1, Michael N. Pollak3, Judy Hirst1* The molecular mode of action of biguanides, including the drug metformin, which is widely used in the treatment of diabetes, is incompletely characterized. Here, we define the inhibitory drug-target interaction(s) of a model biguanide with mammalian respiratory complex I by combining cryo–electron microscopy and enzyme kinetics. We interpret these data to explain the selectivity of biguanide binding to different enzyme states. The primary inhibitory site is in an amphipathic region of the quinone-binding channel, and an additional binding site is in a pocket on the intermembrane-space side of the enzyme. An independent local chaotropic interaction, not previously described for any drug, displaces a portion of a key helix in the membrane domain. Our data provide a structural basis for biguanide action and enable the rational design of medicinal biguanides. T he biguanide metformin is central to the treatment of millions of patients with type 2 diabetes worldwide (1) and has been studied intensely in recent years for treatment of other conditions, including ischemia-reperfusion injury (2, 3), fibrosis (4), viral infections (5), and cancer (6). Optimiza- tion of biguanides for distinctive indications has been hindered by incomplete understanding of their molecular pharmacology, and although preclinical evidence for the antineoplastic action of metformin was sufficient to justify dozens of clinical trials (7), the results have been dis- appointing (8, 9). Biguanides have been reported to target many cellular proteins, including mitochondrial glycerophosphate dehydrogenase (mGPD) (10), presenilin enhancer 2 (11), F1FO adenosine 5′-triphosphate (ATP) synthase (12), cytochrome c oxidase (13), and the chloride intracellular channel 1 (14). Several studies have described biguanide inhibition of mitochon- drial respiratory complex I (proton-translocating NADH:ubiquinone oxidoreductase, where NADH is the reduced form of NAD+) (12, 15–18), sup- porting a mode of biguanide action in which decreased production of ATP from oxidative phosphorylation triggers the activation of adenosine 5′-monophosphate kinase and inhi- bition of adenylate cyclase, leading to bene- ficial downstream effects on gluconeogenic enzymes (in diabetes) and mTOR (in cancer and antiviral treatments) (1, 6, 18, 19). Complex I is a 1-MDa multiprotein assembly that is central to mitochondrial and cellular 1MRC Mitochondrial Biology Unit, University of Cambridge, The Keith Peters Building, Cambridge Biomedical Campus, Cambridge CB2 0XY, UK. 2Structural Biology Laboratory and York Biomedical Research Institute, Department of Chemistry, The University of York, York YO10 5DD, UK. 3Lady Davis Institute of the Jewish General Hospital and Department of Oncology, McGill University, Montreal, QC H3T 1E2, Canada. *Corresponding author. Email: hrb@mrc-mbu.cam.ac.uk (H.R.B.); jh@mrc-mbu.cam.ac.uk (J.H.) metabolism. It oxidizes the NADH produced by oxidation of carbohydrates and lipids to maintain the redox state of the mitochondrial nicotinamide adenine dinucleotide (NAD+) pool, reduces ubiquinone-10 to drive the respiratory chain and oxygen consumption, and pumps protons out of the mitochondrial matrix. This proton pumping contributes to the proton- motive force (PMF) that drives ATP synthesis through oxidative phosphorylation (20). Cryo– electron microscopy (cryo-EM) studies of com- plex I have revolutionized our understanding of its structure, mechanism, and regulation, informing on redox catalysis in the hydro- philic domain, proton translocation across the membrane, and possible mechanisms of coupling between ubiquinone-10 reduction and proton translocation (21, 22). Furthermore, cryo-EM has discriminated different resting states of the enzyme (22–25) on the basis of domain-level reorientations linked to altered conformational states of the quinone-binding channel (Q-channel): the “active” state with a structurally ordered, turnover-ready Q-channel and the pronounced “deactive” state with a locally disordered Q-channel that requires restructuring and reactivation for catalysis. Biguanides bind with an unusual preference to the deactive state of the enzyme (16), but their binding site(s) and modes of interaction are unknown: They are expected to bind in a site downstream of the Fe-S clusters (12) but do not inhibit in a simple competitive manner. Their interaction site is expected to be amphipathic on the basis of the biguanide positive charge and strong correlations between inhibitory po- tency, cytotoxicity, and hydrophobicity (12, 26). We use cryo-EM to reveal the molecular in- teractions of biguanides with mammalian res- piratory complex I, defining how they inhibit catalysis. We identify distinctive binding modes and rationalize biguanide protein-state se- lectivity to enable future implementation of structure-based drug design in the develop- ment of biguanide-based therapies for diverse applications. IM1761092 as a model biguanide for structural studies The antidiabetic biguanides metformin and phenformin are relatively weak inhibitors of complex I, with simple molecular shapes. We sought to avoid technical risks in cryo-EM (excessive adventitious binding and reduced image contrast) from using these compounds in high millimolar concentrations by identify- ing a stronger-binding derivative, which would also exhibit a more distinctive cryo-EM den- sity. For structural investigation of the com- plex I binding site(s) of biguanides, we therefore assessed IM1761092 (hereafter IM1092) (27), a more hydrophobic (logP 2.37) derivative of the metformin-related antidiabetic biguanide phen- formin (logP 0.34) that contains a 3-chloro-4- iodo-phenyl ring (Fig. 1A and fig. S1). IM1092 inhibits cellular oxygen consumption (fig. S1D) and exhibits a stronger inhibition of complex I catalysis in bovine heart mitochondrial mem- branes than phenformin and metformin [half- maximal inhibitory concentration (IC50) in the membrane is more than 10 and 2000 times lower, respectively, depending on the conditions; Fig. 1, B and C]. Its behavior is thus consistent with the reported correlation between biguanide inhibitory potency (IC50) and hydrophobicity (logP) (28). Mammalian mitochondrial membranes “as- prepared” contain a mixture of active and deactive complex I. In the deactive state, an important structural feature of the Q-channel, the loop between TMH1 and -2 in subunit ND3 (ND3 TMH1-2 loop) that carries Cys39 is disordered (24, 25), but in the active state, it is ordered and Cys39 is buried (23, 25, 29, 30). The two states can be discriminated biochem- ically by their sensitivity to N-ethyl maleimide (NEM), which derivatizes ND3-Cys39 in the de- active state, preventing catalysis, but leaves the active state unaffected. We found biguanide inhibition depends on the amount of the de- active state present in membranes (Fig. 1B). As-prepared membranes incubated with both 200 mM IM1092 (10 × IC50) and NEM at 4°C exhibited essentially the same deactive con- tent as biguanide-free controls (fig. S2A), in- dicating that biguanides do not shift the deactive:active population equilibrium in this condition. Furthermore, biguanide inhibition is stronger at higher pH (Fig. 1C). As biguanides [pKa ~11 (where Ka is the acid dissociation constant) (31)] remain singly protonated at all pH levels tested, the pH dependence likely arises from changes in the protein, such as in local charges on residue side chains or phospholipid headgroups, or conformational changes. Hav- ing established that IM1092 inhibits cellular Bridges et al., Science 379, 351–357 (2023) 27 January 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E oxygen consumption and shares the same se- lectivity for inhibiting catalysis by binding to the deactive state of complex I that is observed for antidiabetic biguanides, as well as the same pH-dependent mode of action, we proceeded with this tighter-binding synthetic biguanide as a representative model compound for struc- tural studies for the technical reasons outlined previously. To investigate the selectivity of biguanides for different states, IM1092 was added to the mixed population of states in the purified rest- ing enzyme, without purposeful deactivation, for cryo-EM. No substrates were added be- cause inhibition is weaker when biguanides are added during catalysis, rather than before (12). IM1092 has an IC50 value of 86 mM for catalysis of detergent-solubilized, purified bo- vine complex I (fig. S2B), and cryo-EM grid conditions with 350 mM IM1092 were chosen to maximize binding within the limits of pH- and biguanide-induced protein aggregation at high concentrations (fig. S2). When purified complex I (fig. S3) was incubated in the same conditions as for cryo-EM grid preparation and then diluted into inhibitor-free assay buffer, 98% of the control catalytic rate was recovered (fig. S2F), demonstrating full reversibility of inhi- bition under this condition. Overview of cryo-EM particle populations A total of 17,203 micrographs collected from a single cryo-EM grid yielded 598,287 good par- ticles, which refined to an estimated global res- olution of 2.1 Å (fig. S4), but with such a high degree of heterogeneity that it was not possible to model parts of the map. Subsequent global classification yielded three major classes resem- bling the active and deactive states mentioned A H2N N N NH2 NH2 Metformin H2N N H N NH2 NH2 Phenformin H2N N H N NH2 NH2 IM1092 Metformin 99 mM 52 mM 33 mM B 120 100 ) M m ( 0 5 C I 80 60 40 20 5 4 3 2 1 Phenformin 3.8 mM 539 µM 131 µM I Cl 0 0 50 100 Deactive content (%) 0 50 0 Deactive content (%) 100 C ) M m ( 0 5 C I 300 250 200 150 100 50 0 Metformin M m 8 3 M m 8 7 1 M m 1 2 1 M m 5 1 M m 8 M m 9 6 7 8 pH 9 9D 9D(i) 6 5 4 3 2 1 0 Phenformin M m 3 . 5 M m 4 . 1 M µ 0 3 4 M µ 1 9 1 M µ 3 4 M µ 0 5 6 7 8 9 9D 9D(i) Hp ) M µ ( 0 5 C I 50 40 30 20 10 0 M µ 3 4 M µ 7 2 M µ 4 M µ 1 2 M n 6 4 8 M n 6 1 8 6 7 8 Hp 9 9D 9D(i) 50 40 30 20 10 ) M µ ( 0 5 C I IM1092 32 µM 20 µM 4 µM 0 0 50 100 Deactive content (%) IM1092 Fig. 1. Characterization of biguanide effects on catalysis and regions of structural interest. (A) Chemical structures of metformin, phenformin, and IM1092 in monoprotonated form. (B) Correlation between membrane deactive complex I content and IC50 for metformin (gray), phenformin (teal), and IM1092 (orchid). Error bars represent SEM for deactive content and 95% confidence intervals for IC50. Data are fit to an exponential regression for visualization. (C) Effect of pH on IC50 in bovine heart membranes for metformin (gray), phenformin (teal), and IM1092 (orchid). 9D, pH 9 deactivated membranes; 9D(i), pH 9 deactivated membranes measured in the presence of antimycin A to inhibit complex III and the alternative oxidase to oxidize quinol by a different route, confirming inhibition is on complex I. Error bars represent 95% confidence intervals. Table 1. Key characteristics of classes presented in this work. Unclear, unclear binding position, but confident identity; unidentified, the presence of density of unknown nonprotein origin. Model Classification scheme PDB ID Features Q-channel (C-Cmask) ND5 and NDUFB4 interface ND2, NDUFB5, and NDUFA11 (C-Cmask) Active inhibitor-free ............................................................................................................................................................................................................................................................................................................................................ Active-1092-i ............................................................................................................................................................................................................................................................................................................................................ Active-1092-ii ............................................................................................................................................................................................................................................................................................................................................ Active-1092-iii ............................................................................................................................................................................................................................................................................................................................................ Active-1092-iv ............................................................................................................................................................................................................................................................................................................................................ Deactive-1092-i ............................................................................................................................................................................................................................................................................................................................................ Deactive-1092-ii ............................................................................................................................................................................................................................................................................................................................................ Deactive-1092-iii ............................................................................................................................................................................................................................................................................................................................................ Deactive-1092-iv ............................................................................................................................................................................................................................................................................................................................................ Deactive-1092-v ............................................................................................................................................................................................................................................................................................................................................ Deactive-1092-vi ............................................................................................................................................................................................................................................................................................................................................ Slack-1092-i ............................................................................................................................................................................................................................................................................................................................................ Slack-1092-ii ............................................................................................................................................................................................................................................................................................................................................ Inhibitor-free (fig. S5) Q-channel (fig. S7) Lateral helix (fig. S8) Lateral helix (fig. S8) Lateral helix (fig. S8) Q-channel (fig. S7) Q-channel (fig. S7) Q-channel (fig. S7) Lateral helix (fig. S8) Lateral helix (fig. S8) Lateral helix (fig. S8) Q-channel (fig. S7) Q-channel (fig. S7) NDUFC2 N terminus IM1092-bound (0.49) IM1092-bound (0.52) Unidentified IM1092-bound (0.63) IM1092-bound (0.60) Unidentified IM1092-bound (0.68) IM1092-bound (0.65) IM1092-bound (0.64) IM1092-bound (0.67) IM1092-bound (0.65) IM1092-bound (0.64) No density Unidentified Unidentified Unidentified Unidentified IM1092-bound (0.66) IM1092-bound (0.72) IM1092-bound (0.68) IM1092-bound (0.66) IM1092-bound (0.71) IM1092 unclear IM1092 unclear IM1092-bound (0.64) Intact Mixed Intact Displaced Disordered Mixed Mixed Mixed Intact Displaced Disordered Mixed Mixed 7QSD 7R41 7R42 7R43 7R44 7R45 7R46 7R47 7R48 7R4C 7R4D 7R4F 7R4G Bridges et al., Science 379, 351–357 (2023) 27 January 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E above and a state called slack (supplementary text), which is of unknown functional relevance but has been described previously in cryo-EM studies of bovine complex I (22, 24, 25). Inhibitor- free reference maps of bovine complex I from separate inhibitor-free preparations in deter- gent [figs. S5 and S6 and EMD-3731 (24)] were used for comparison to identify and evaluate features found in the maps with IM1092 present. Three regions of interest are (i) densities occupy- ing the Q-channel; (ii) unusually poor density in a portion of the C-terminal lateral helix in sub- unit ND5 and the adjacent subunit NDUFB4; and (iii) altered density at the expected po- sition of the NDUFC2 subunit N terminus. Different classification strategies were tested to disentangle the different states and IM1092 interactions, leading to implementation of a “local-first” classification regime (supplemen- tary text) used in two separate schemes to describe 12 distinct classes (Table 1 and figs. S7 to S9). Biguanide-binding site 1: Ubiquinone-binding channel Density was observed in the Q-channel, close to its entrance from the membrane, in all three major classes. Using local-first classification (fig. S7), we separated the particles into one active class (Active-1092-i; fig. S10), three de- active classes (Deactive-1092-i, -ii, and -iii; figs. S11 to S13), and two slack classes (Slack-1092-i and -ii; Table 1 and figs. S14 and S15). The ma- jor classes were assigned by global comparisons to reference maps and key local features (table S1 and supplementary text). Densities for IM1092 in the Q-channel are clear in the Deactive-1092-i, -ii, and -iii and Slack-1092-ii maps (Fig. 2, C to E and H, and figs. S16 and S17), and the cross- correlation fits (C-Cmask) (32) for the fit of the IM1092 molecule into its density were high (0.66 to 0.72) (Table 1). In Slack-1092-i, although the density is consistent for the chloro-iodo-phenyl moiety, the biguanide moiety of the putative IM1092 molecule was insufficiently resolved to confidently model its orientation (Fig. 2G). The density in the Active-1092-i map occupies a position similar to that of the densities ob- served in the other classes (Fig. 2F), but a smaller additional density feature is also observed fur- ther into the Q-channel so that, together, they resemble density observed in the Q-channel of the active-apo (inhibitor-free) class of bovine complex I in nanodiscs (EMD-14133) (22). With an IM1092 molecule refined in different orientations into the density in the Active- 1092-i map near the Q-channel entrance, C- Cmask values are low (<0.5), and the shape of the density was visibly a poor fit; the second density was too small and featureless to as- certain its origin, and the identity of these two densities remain unconfirmed. Overall, ~45 to ~60% of the total population and ~56 to ~75% of the (deactive and slack) population presents clear evidence of IM1092 occupying the Q-channel. In the deactive and slack states, IM1092 binds in an amphipathic site straddling two zones of the Q-channel: the hydrophobic region next to the exit (Fig. 2A) and the charged central re- gion of the channel (33). The chloro-iodo-phenyl group of IM1092 points toward the channel exit (Fig. 2A), forming weak halogen bonds from the Cl to the ND1-Pro48 carbonyl and from the I to the S of ND1-Met225, as well as van der Waals interactions with ND1-Phe224, ND1-Phe220, ND1-Leu55, and NDUFS7-Trp46. The halogen bonds are specific to IM1092, rel- ative to phenformin, likely contributing to its higher binding affinity; and metformin, which lacks the phenyl ring, is not stabilized by the above interactions or by p-stacking to ND1- Phe224, consistent with its much lower binding affinity. The biguanide moiety faces into the channel, toward the charged region, and adopts a range of different orientations (Fig. 2, C to E and H). In the Deactive-1092-i and Slack- 1092-ii states, it forms a cation-p interaction with NDUFS7-Trp46 and, in the Deactive-1092-i state, a weak ionic interaction with the ND1- Glu24 carboxyl also. In the Deactive-1092-ii and -iii states, its orientation brings it closer A 90° B Deactive ND1 Glu-24 Slack Thr-21 Arg-77 Tyr-228 Arg-77 Phe-224 IM1092 Phe-224 Active Trp-46 IM1092 NDUFS7 ND1 Pro-48 cluster N2 NDUFS7 IM1092 ND1 Met-225 NDUFS7 C 4.4 Å 4.4 Å Deactive-1092-i F D G 3.1 Å 2.5 Å 4.5 Å E 3.3 Å 2.8 Å 3.5 Å 4.8 Å Deactive-1092-ii Deactive-1092-iii H 4.6 Å 4.9 Å Active-1092-i Slack-1092-i Slack-1092-ii Fig. 2. Binding of IM1092 in the Q-channel. (A) Overview showing the location of the biguanide-binding site and inset showing a closer view and bonding interactions of the chloro-iodo-phenyl group for Deactive-1092-i. Orange arrow shows the route for exit from the Q-channel into the lipid bilayer. (B) Overlay of models for the Active-1092-i; Deactive-1092-i, -ii, and -iii; and Slack-1092-i and -ii states, aligned to subunit ND1, showing the location and variability of biguanide-binding orientations and relative position of NDUFS7-Arg77. NDUFS7-Arg77 and ND1-Phe224 are white in the active models, mint or orchid in the slack models, and black in the deactive models. (C to H) Cryo-EM difference map densities (composite versus models) for biguanides bound to Deactive-1092-I (C), Deactive-1092-ii (D), Deactive-1092-iii (E), Active-1092-i (F), Slack-1092-i (G), and Slack-1092-ii (H). The insets show the difference map density for the biguanide for each model shown. Biguanides are not modeled in (F) and (G), owing to uncertainties in the ligand identity or orientation. Side-chain and ligand density for (C) to (H) are shown in fig. S16. Bridges et al., Science 379, 351–357 (2023) 27 January 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E to ND1-Glu24, and it forms a hydrogen bond with the ND1-Tyr228 hydroxyl. In each state with modeled biguanide, the experimental IM1092 map density at higher thresholds is consistent with a mixture of inhibitor binding poses. Notably, the conformation of the biguanide- binding region of the Q-channel differs between the three major states (21–23), regardless of whether an inhibitor is bound (Fig. 2B). No- tably, NDUFS7-Arg77 swings in an arc from its position in the active state, with its side chain pointing toward the Q-channel exit, to point approximately toward cluster N2 in the de- active state, adopting an intermediate position in the slack state (Fig. 2B). The NDUFS7-Arg77 guanidinium is ~7 Å (slack) and 10 to 14 Å (de- active) away from the bound biguanide (N-N distance), but only ~4 Å away when IM1092 is refined into the density in the Active-1092-i map in the same orientation, indicating that repulsion between their positive charges may disfavor biguanide binding in the active state. Available Protein Data Bank (PDB) models for mammalian, fungal, bacterial, and plant complexes I were compared to assess the con- servation of key residues interacting with the biguanide moiety; NDUFS7-Trp46, ND1-Glu24, ND1-Tyr228, and NDUFS7-Arg77 are conserved in all species surveyed. The only high-quality reported human mutation in these residues in the ClinVar database (34) is the ND1-E24K (Glu24→Lys) mutation associated with Leber hereditary optic neuropathy–mitochondrial en- cephalopathy, lactic acidosis, and stroke-like episodes (LHON-MELAS) overlap syndrome (35). Human mutations of the residues involved in the biguanide-binding site are therefore rare, consistent with their important role in enzyme function. Because they are mitochondrially en- coded and/or essential for function, these re- sidues cannot easily be artificially mutated in a mammalian system. Methods for mutating mitochondrial DNA are advancing, but site spe- cificity and the efficiency of reaching homoplas- mic edits remain challenging (36). Biguanide interaction site 2: ND5 lateral helix and NDUFB4 interface Transmembrane subunit ND5 contains an un- usual long helix that runs laterally alongside ND4 and ND2, which has been proposed to either stabilize the proton-pumping modules or act as a transmissive element in proton pumping (37). Substantial evidence of disorder at the ND5 lateral helix–subunit NDUFB4 in- terface was observed in preliminary maps, so the dataset was subject to a separate local-first classification that focused on this region (figs. S8 and S18). The strategy yielded three major classes: one with a typical well-ordered ND5 lateral helix and NDUFB4 and two with dis- tortion or disordering of ND5 residues 547 to 564 and nearby NDUFB4 residues 77 to 92. The classes were further separated into Active- A NDUFB8 Pro-77 Lys-564 Tyr-288 ND4 Lys-218 Tyr-148 3.0 Å His-213 2.6 Å 3.1 Å 2.6 Å 75° Asp-554 Glu-559 ND5 Lys-547 45° Phe-92 NDUFB4 ND5 Trp-557 2.7 Å NDUFA11 NDUFB4 Thr-85 NDUFA11 B Intact: Active-1092-ii C Displaced: Active-1092-iii D Disordered: Active-1092-iv Fig. 3. Biguanide-induced distortion and disordering of the ND5 lateral helix and NDUFB4 loop. (A) Location of the structure disturbance and two views of the model of Deactive-1092-iv with hydrogen- bonding interactions indicated with black dotted lines and distances indicated. Orchid, NDUFB4; teal, ND5; orange, NDUFA11; dark gray, NDUFB8. Side-chain density for this region of Deactive-1092-iv is shown in fig. S18. (B to D) Models and composite cryo-EM maps for Active-1092-ii (B), Active-1092-iii (C), and Active-1092-iv (D) showing progressive disordering within the series. Details of p-bulge stabilizing interactions and equivalent disordering for the deactive states (Deactive-1092-iv, -v, and -vi) are shown in fig. S18. 1092-ii, -iii, and -iv (figs. S19 to S21); Deactive- 1092-iv, -v, and -vi (Table 1 and figs. S22 to S24); and three slack classes, which all displayed poor density for the downstream ND5 lateral helix and were not further investigated. Overall, ~65% of the protein population was perturbed in this region, with similar proportions observed for active, deactive, and slack (~52, ~66, and ~58%, respectively). Deactive classes with a disrupted helix exhibit a small “opening” of the angles be- tween the membrane and hydrophilic domains and distal and proximal membrane domains compared with their better-ordered equivalents (fig. S25 and supplementary text). This interesting region of the usually well- ordered ND5 lateral helix does not form a perfect a helix even in inhibitor-free active or deactive mammalian enzyme structures (21–25). The Active-1092-ii and Deactive-1092-iv mod- els match the well-ordered inhibitor-free active model in this region. Two p-bulges (fig. S18) are stabilized by interactions between the lateral helix and nearby waters and by ionic and hy- drogen bonding to ND4 (Fig. 3A, right). In Active-1092-iii and Deactive-1092-v (re- ferred to as displaced), a short portion of the ND5 lateral helix is altered: Lys547 to Ser550 are disordered, and an interruption of the helical structure at Leu562 to Pro563 allows a short stretch of helix (residues 550 to 559) to move outwards, away from the complex, and later- ally, along the membrane plane (Fig. 3 and fig. S18). A loop in NDUFB4 at residues 78 to 83, which usually wraps around the lateral helix, becomes disordered from Pro77 to Leu91. In Active-1092-iv and Deactive-1092-vi (referred to as disordered), the whole ND5 region from Lys547 to Lys564 appears disordered, along with the NDUFB4 loop described previously (Fig. 3D and fig. S18). Although not observed in the detergent-solubilized protein, the bovine enzyme in nanodiscs (22) contains two phospholipids near to the distorted region. No nearby density features can be interpreted as a tightly bound biguanide. Considering the positive charge on IM1092, it may interact with ND5-Asp554 and/or Glu559, thereby disrupting hydrogen bonding from the lateral helix to subunit ND4, or interact with nearby stabilizing phospholipids. Gen- erally, p-bulges are energetically unfavorable elements (38) that require stabilization by hy- drogen bonding to polar side chains or water molecules (39). The strained nature of this region may make it particularly prone to destabilization by guanidium-like biguanides. No mutations of Asp554 or Glu559 are observed in the ClinVar database (34), and acidic residues in these posi- tions are conserved in current mammalian and plant structures, which suggests their important role in stabilizing the membrane domain. Biguanide-binding site 3: ND2, NDUFB5, and NDUFA11 Density that matches IM1092 was observed in a pocket formed by subunits ND2, NDUFB5, and NDUFA11 on the intermembrane-space Bridges et al., Science 379, 351–357 (2023) 27 January 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E A C 5BFUDN IM1092 ND2 B D NDUFC2 ND2 Glu-347 Asn-197 Trp-122 IM1092 Phe-123 Val-140-C NDUFB5 NDUFA11 5BFUDN ND2 Fig. 4. Location of the biguanide binding at the ND2-NDUFB5-NDUFA11 interface. (A) Overall location of the binding site. Purple, NDUFC2; teal, ND2; orange, NDUFB5. (B) Overlay of all models aligned to ND2. White, NDUFA11; Val-140-C, C terminus of NDUFA11. (C) Surface representation (Active-1092-ii model) showing subunit NDUFC2 in cartoon, with atoms shown for IM1092. (D) Same view from the active inhibitor-free model showing the atoms from residues 1 to 7 of NDUFC2. Local interactions and density for the individual maps and models are shown in figs. S26 and S27. side of the enzyme (Fig. 4 and figs. S26 and S27). The pocket is occupied by the N-terminal seven residues of NDUFC2 in the inhibitor-free en- zyme, both in n-dodecyl-B-D-maltoside (DDM) and in all states of the bovine enzyme in nano- discs (22). The NDUFC2 N terminus is displaced by the biguanide, with inhibitor densities ob- served in most classes in this study (figs. S26 and S27). The biguanide moiety is stabilized by an ionic interaction with ND2-Glu347 and by hydrogen bonds with the ND2-Tyr196 back- bone carbonyl, ND2-Thr199 hydroxyl, ND2-Asn197 side chain, and NDUFA11-Val140 C-terminal car- boxyl, as well as nearby water molecules re- solved in some maps (figs. S26 and S27). The IM1092 bound in this site does not interfere with any known catalytically relevant struc- tural elements in the complex and so is likely to represent a noninhibitory interaction. Independence of interaction sites The two local classification schemes used (figs. S7 and S8) each yielded six models, and a summary of their key features is shown in Table 1. All six classes from Q-channel clas- sification (fig. S7) had poor density for the ND5-NDUFB4 interface, consistent with a mixture of the three lateral helix states being represented there. Overall, ~65% of the im- aged protein population is disrupted in this region (fig. S8), and disorder is observed in all three major classes, as well as in states with (e.g., Deactive-1092-i) or without clear binding of IM1092 at the Q-channel (Active-1092-i). Fur- thermore, in the classes from the lateral helix classification (fig. S8), density for IM1092 was observed in the Q-channel of deactive classes with both ordered and displaced lateral helix. Taken together, biguanide binding in the Q- channel and the state of the lateral helix are not correlated. IM1092 is observed in the ND2, NDUFB5, and NDUFA11 pocket regardless of the occupancy of the Q-channel or the status of the lateral helix (Table 1). Therefore, the three interaction sites are independent of one other. Discussion Biguanide access to complex I binding sites Substantial inhibition of complex I in vivo re- quires the biguanide positive charge, an in- herent feature of the two cojoined guanidinium moieties at physiological pH (31), to drive biguanide accumulation in the mitochondrial matrix (40) by up to 1000-fold relative to the cytosol in response to the mitochondrial PMF. This mitochondrial concentrating effect makes the intramitochondrial biguanide concentra- tion sufficient for the inhibition of targets with only relatively weak biguanide affinity, such as complex I, making them as potentially rele- vant as targets outside of the mitochondrial matrix that exhibit greater intrinsic affinities, such as rat mGPD (10). Recent work has ques- tioned mGPD itself as a therapeutic target be- cause metformin was found to be noninhibitory against the human enzyme (41). All three inter- action or binding sites described in complex I contain acidic residues close to the membrane- aqueous interface, suggesting that IM1092 ac- cesses them from the membrane, likely with the chloro-iodo-phenyl group acting as an an- chor into the hydrophobic membrane core and the hydrophilic biguanide moiety inter- acting with the negatively charged phosphate headgroups, as proposed previously (42, 43). The most likely route of access for hydrophobic biguanides to the Q-channel is therefore via the matrix-facing phospholipid leaflet. Although the Q-channel is reproducibly well ordered in the active state, deactive (or slack) states have mobile regions of the Q-channel loops that face the mitochondrial matrix (21–24, 44), so it is also possible that the Q-channel may be- come exposed to the matrix in these states, providing an alternative route for hydrophilic biguanides with poor membrane solubility, such as metformin, to enter the Q-channel. Major inhibitory site and selectivity for the deactive state Biguanides, including IM1092 and the anti- diabetic compounds metformin and phenformin, all share a preference to bind to the deactive state of complex I. This behavior is not observed for any other class of complex I inhibitor, and we expect all three biguanides to share the same inhibitory binding site and mode of ac- tion. In comparison with canonical hydropho- bic complex I inhibitors such as rotenone, biguanides are relatively hydrophilic, water- soluble molecules that are unlikely to bind in highly hydrophobic or membrane-intrinsic sites. We suggested previously that metformin might interact at the junction of the hydro- philic and hydrophobic domains where a set of mobile elements (in subunits NDUFS7, NDUFS2, ND3, and ND1) change their conformation be- tween the active and deactive states by bind- ing to a resting state and preventing a return to catalysis (12). Our structures now demon- strate that the biguanide-binding site is inside the Q-channel, in a region that likely becomes exposed to the matrix in the deactive state, and adjacent to the mobile element in NDUFS7 that carries Arg77 and that switches its confor- mation between the active and deactive states. In the Q-channel, IM1092 binds in an am- phipathic region with the biguanide moiety stabilized by hydrogen bonding, cation-p inter- actions, and ionic bonding, and the hydropho- bic chloro-iodo-phenyl group is stabilized by weak halogen bonding and van der Waals in- teractions. These additional stabilizing inter- actions between the hydrophobic portion of the Q-binding site and hydrophobic biguanides explain the relationship observed previously between biguanide hydrophobicity and inhib- itory potency (12). Other neutral and highly hydrophobic ligands (DDM, cholate, rotenone, and IACS-2858) have also been observed binding Bridges et al., Science 379, 351–357 (2023) 27 January 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E in overlapping sites in various protein states (fig. S28) (21, 22, 45). Yet biguanides remain distinctive inhibitors in their state selectivity, shown here and previously (16). Of the other inhibitors, only rotenone has been demon- strated to bind to a deactive-like open state, but the binding was not state specific (21) and the protein interactions and in vivo effects of biguanides and rotenone are different (46, 47). Despite the locations of the biguanide- and ubiquinone-binding sites overlapping, met- formin does not display classical competitive behavior (12). This is likely because neutral, hydrophobic inhibitors such as piericidin A, IACS-2858, and acetogenin (45, 48, 49) com- pete for (active-like) enzyme states capable of forming the Michaelis complex, but biguanide binding occurs most readily in deactive-like states that are not preorganized, owing to dis- ordering of the Q-channel loops in NDUFS2, ND3, and ND1 (24). The deactive state of com- plex I is stabilized by high pH (50), and we see a strong correlation between pH and inhibi- tory potency. Weak biguanide inhibition of active complex I could originate either from binding to the Q-channel or from an inhib- itory nature to the ND5 lateral helix interac- tion. The conformation of the NDUFS7 b1–b2 loop that contains Arg77 differs between the active, deactive, and slack states. Biguanides bind 7 to 14 Å from Arg77 in deactive and slack states, but the arginine guanidinium position is shifted relative to the positions in the DDM- and cholate-bound bovine nanodisc deactive and slack states (PDB IDs 7QSM and 7QSO) (fig. S28) (22). In all active models, Arg77 is much closer to the putative biguanide-binding site (Fig. 2B), presenting a source of steric hin- drance and charge repulsion that acts against strong inhibitory binding and explains the preference of biguanide inhibitors to bind to NEM-sensitive (deactive) complex I [manifest- ing as a lower IC50 for deactivated complex I (Fig. 1B) (16)]. Biguanides are less powerful inhibitors when added during catalysis rather than before (12), and metformin has been pro- posed to act by slowing down activation (51) rather than by classical inhibition. Our data support an inhibitory mode that primarily acts by preventing reactivation of the resting de- active state, and we conclude that the major inhibitory interaction site of biguanides is the Q-channel. Local chaotropic drug-protein interactions A key feature of biguanide interactions with complex I is the displacement and disordering of a portion of the ND5 lateral helix. This type of localized disruption has not been observed previously but may be facilitated by nonspe- cific interactions with phospholipids (42). Considering the similarity of biguanides to the well-known chaotrope and protein denaturant guanidinium, the biguanide may be attracted to the negatively charged ND5-Asp554 and ND5-Glu559 and specifically destabilize the hydrogen-bonding networks between the lat- eral helix and subunit ND4. In particular, biguanide binding could weaken the inter- actions of the lateral helix p-bulge segment, allowing secondary structure shifts to form an a helix and a disordered loop. After this process, the unraveled segment of the helix would no longer provide lateral support to keep the strained junction between ND2 and ND4 (52) tightly together, explaining the slight “opening” of the proximal and distal mem- brane domain interface (fig. S25). Protein sta- bility assays (fig. S2) suggest that complex I is stabilized close to the IM1092 IC50 ranges but is destabilized at higher concentrations, and we interpret this behavior to indicate that bind- ing to the Q-channel (and/or ND2-NDUFB5 in- terface) stabilizes the protein and that chaotropic action occurs at higher concentrations. Our structures in this study are in a detergent micelle with tightly bound lipids present, and the chao- tropic interaction site we observe is within the phospholipid headgroup plane. Notably, any inhibitory consequences of structural distur- bance in this study are fully reversible, which is demonstrated by full recovery of activity after inhibitor dilution (fig. S2F). We propose that biguanides such as metformin, phenformin, and related drug leads could exert similar local chaotropic actions on any number of cellular proteins, especially membrane proteins, to in- hibit or stimulate their usual functions, present- ing a novel mode of enzyme-drug interaction and a potential explanation for the breadth of biguanide targets identified thus far. Implications for in vivo mechanism of action Although other complex I inhibitor classes have been proposed as possible therapeutic com- pounds (53–56), biguanides appear to offer a lower toxicity profile than neutral species such as rotenone, with reduced risk of Parkinsonism (47). The key factor may lie in self-limitation of action in the form of negative feedback at the organelle level, meaning that their mechanism of action lowers the risk of complete inhibi- tion of the respiratory chain. The membrane potential leads to increased concentration of biguanides at the complex I site of action, but as complex I is inhibited, the membrane potential falls (57), creating an equilibrium that limits the concentration and therefore the degree of inhibition. Toxicity may also be in- fluenced by the ready reversibility of biguanide inhibition seen here and previously (12), com- pared with the irreversibility of very hydropho- bic, neutral compounds. In vivo, the active-like state that binds canonical inhibitors is expected to be prevalent in normal tissues (44), where the opportunities for biguanides to inhibit may be restricted to the weak binding observed in this study against active-like states, or to the binding of putative catalytic intermediates where NDUFS7-Arg77 has moved away from the inhibitory binding site. The deactive state of complex I forms during oxygen starvation (51), such as during ischemia (2) and in solid tumor microenvironments (58). Discrimina- tion of biguanides for this state means that, unlike canonical inhibitors, hypoxic tissues can be selectively targeted with less risk of also compromising mitochondrial respira- tion in tissues with normally functioning (active) complex I. In such low-oxygen environments, only complex I in the active state may work in the reverse direction, using the PMF and ubiquinol to drive NAD+ reduction and reac- tive oxygen species production (51). By targeting the deactive state formed in these environ- ments, preventing reactivation, biguanides may diminish reactive oxygen species production by reverse electron transfer, as has been ob- served previously for metformin (59). Metformin has proven to be a safe and ef- fective biguanide for diabetes treatment, and our data now provide a structure-based ratio- nale for its mode of action on complex I. This work also offers a basis for future structure-based biguanide drug design for the state-specific inhibition of complex I for other potential ther- apeutic applications (such as cancer treatment) in which metformin has been less successful, as well as improved prediction of additional protein target binding sites for biguanide drugs. REFERENCES AND NOTES 1. M. Foretz, B. Guigas, L. Bertrand, M. Pollak, B. Viollet, Cell Metab. 20, 953–966 (2014). 2. K. Skemiene et al., Biomolecules 10, 1400 (2020). 3. X. Wang et al., PLOS ONE 12, e0182777 (2017). 4. N. Sato et al., Respir. Res. 17, 107 (2016). 5. S. Lehrer, World Acad. Sci. J. 2, 1 (2020). 6. H. Zhao, K. D. Swanson, B. Zheng, Trends Cancer 7, 714–730 (2021). 7. M. Pollak, Nat. Med. 20, 591–593 (2014). 8. S. Kordes et al., Lancet Oncol. 16, 839–847 (2015). 9. P. J. Goodwin et al., J. Clin. Oncol. 31, 1033–1033 (2013). 10. A. K. Madiraju et al., Nature 510, 542–546 (2014). 11. T. Ma et al., Nature 603, 159–165 (2022). 12. H. R. Bridges, A. J. Y. Jones, M. N. Pollak, J. Hirst, Biochem. J. 462, 475–487 (2014). 13. T. E. LaMoia et al., Proc. Natl. Acad. Sci. U.S.A. 119, e2122287119 (2022). 14. F. Barbieri et al., Front. Oncol. 9, 135 (2019). 15. M. Y. El-Mir et al., J. Biol. Chem. 275, 223–228 (2000). 16. S. Matsuzaki, K. M. Humphries, Biochemistry 54, 2011–2021 (2015). 17. M. R. Owen, E. Doran, A. P. Halestrap, Biochem. J. 348, 607–614 (2000). 18. G. Rena, D. G. Hardie, E. R. Pearson, Diabetologia 60, 1577–1585 (2017). 19. R. A. Miller et al., Nature 494, 256–260 (2013). 20. J. Hirst, Annu. Rev. Biochem. 82, 551–575 (2013). 21. D. Kampjut, L. A. Sazanov, Science 370, eabc4209 (2020). 22. I. Chung et al., Nat. Commun. 13, 2758 (2022). 23. A. A. Agip et al., Nat. Struct. Mol. Biol. 25, 548–556 (2018). 24. J. N. Blaza, K. R. Vinothkumar, J. Hirst, Structure 26, 312–319. e3 (2018). 25. J. Zhu, K. R. Vinothkumar, J. Hirst, Nature 536, 354–358 (2016). 26. T. Sakai et al., Sci. Rep. 11, 4852 (2021). 27. S. W. Kim et al., “Guanidine Compounds and use thereof,” US patent application no. 2017/0073331 A1 (16 March 2017). 28. H. R. Bridges, V. A. Sirviö, A. N. A. Agip, J. Hirst, BMC Biol. 14, 65 (2016). Bridges et al., Science 379, 351–357 (2023) 27 January 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E 29. M. Babot, A. Birch, P. Labarbuta, A. Galkin, Biochim. Biophys. Acta 1837, 1083–1092 (2014). 30. E. V. Gavrikova, A. D. Vinogradov, FEBS Lett. 455, 36–40 57. Z. Drahota et al., Physiol. Res. 63, 1–11 (2014). 58. A. Emami Nejad et al., Cancer Cell Int. 21, 62 (2021). 59. C. Batandier et al., J. Bioenerg. Biomembr. 38, 33–42 (2006). (1999). 31. D. Kathuria, A. A. Bankar, P. V. Bharatam, J. Mol. Struct. 1152, ACKN OWLED GMEN TS 61–78 (2018). 32. P. V. Afonine et al., Acta Crystallogr. D Struct. Biol. 74, 814–840 (2018). 33. J. G. Fedor, A. J. Y. Jones, A. Di Luca, V. R. I. Kaila, J. Hirst, Proc. Natl. Acad. Sci. U.S.A. 114, 12737–12742 (2017). 34. M. J. Landrum et al., Nucleic Acids Res. 46, D1062–D1067 (2018). 35. E. L. Blakely et al., Eur. J. Hum. Genet. 13, 623–627 (2005). 36. P. Silva-Pinheiro, M. Minczuk, Nat. Rev. Genet. 23, 199–214 (2022). 37. R. G. Efremov, R. Baradaran, L. A. Sazanov, Nature 465, 441–445 (2010). 38. J. Ludwiczak et al., Sci. Rep. 9, 6888 (2019). 39. J. P. Cartailler, H. Luecke, Structure 12, 133–144 (2004). 40. F. Davidoff, J. Biol. Chem. 246, 4017–4027 (1971). 41. M. J. MacDonald, I. H. Ansari, M. J. Longacre, S. W. Stoker, Diabetes 70, 1575–1580 (2021). 42. G. Schäfer, Biochem. Pharmacol. 25, 2005–2014 (1976). 43. N. Samart, C. N. Beuning, K. J. Haller, C. D. Rithner, D. C. Crans, Langmuir 30, 8697–8706 (2014). 44. Z. Yin et al., Nat. Commun. 12, 707 (2021). 45. I. Chung et al., Sci. Adv. 7, eabg4000 (2021). 46. M. Konopleva et al., Targeting Oxidative Phosphorylation with a Mitochondrial Complex I Inhibitor is limited by Mechanism-based Toxicity, 1506700/v1, Research Square [Preprint] (2022). 47. S. Heinz et al., Arch. Toxicol. 93, 2603–2615 (2019). 48. H. R. Bridges et al., Nat. Commun. 11, 5261 (2020). 49. D. N. Grba et al., J. Biol. Chem. 298, 101602 (2022). 50. A. B. Kotlyar, V. D. Sled, A. D. Vinogradov, Biochim. Biophys. Acta 1098, 144–150 (1992). 51. A. Galkin, S. Moncada, Interface Focus 7, 20160104 (2017). 52. A. Di Luca, V. R. I. Kaila, Biochim. Biophys. Acta Bioenerg. 1862, 148382 (2021). 53. N. Madhusudhan et al., ACS Chem. Biol. 15, 158–170 (2020). 54. J. R. Molina et al., Nat. Med. 24, 1036–1046 (2018). 55. A. Naguib et al., Cell Rep. 23, 58–67 (2018). 56. D. A. Russell et al., J. Nat. Prod. 83, 1829–1845 (2020). We thank D. Chirgadze (Cambridge), J. Radecke, and Z. Yang (eBIC) for their assistance with grid screening and data collection; Diamond for access and support of the cryo-EM facilities at the UK national electron Bio-Imaging Centre (eBIC), proposals BI22238-10 and EM17057-23, funded by the Wellcome Trust, MRC, and BBSRC; members of the Hirst laboratory, especially N. Agip, for their participation in bovine mitochondrial preparations; S. McLaughlin (LMB, Cambridge) for access to the nano-DSF; T. Terwilliger (Los Alamos, NM) for guidance with map sharpening and composite map generation; S. Ding (MBU) for recording mass spectrometry data; H. Prag (MBU) for recording liquid chromatography data; P. Gierth (Cambridge) for recording nuclear magnetic resonance data; J. Bosak (MBU) for culturing and authenticating cells; and M. Iadanza (Leeds) and P. Afanasyev (ETH, Zurich) for their scripts deposited on Github. Funding: This work was supported by the Medical Research Council (MC_U105663141 and MC_UU_00015/2 to J.H.) and by ImmunoMet Therapeutics Inc. Author contributions: Conceptualization: H.R.B., M.N.P., and J.H. Methodology: H.R.B. and J.N.B. Validation: H.R.B. Formal analysis: H.R.B. Investigation: H.R.B., J.N.B., Z.Y., and I.C. Data curation: H.R.B. Writing – original draft: H.R.B. Writing – review and editing: H.R.B., J.N.B., Z.Y., I.C., M.N.P., and J.H. Visualization: H.R.B. Supervision: J.H. Project administration: H.R.B. and J.H. Funding acquisition: J.H. Competing interests: ImmunoMet Therapeutics Inc. are the inventors on patent US2017/007331 for the biguanide compound IM1761092 used in this study. M.N.P. is on the scientific advisory board of ImmunoMet Therapeutics and holds shares in the company. The authors declare that they have no further competing interests. Data and materials availability: The structure data accession codes are EMD-14251, EMD-14252, EMD- 14253, EMD-15254, EMD-15355, and PDB-7R41 (Active-1092-i); EMD-4256, EMD-14257, EMD-14258, EMD-14259, EMD-14260, and PDB-7R42 (Active-1092-ii); EMD-14261, EMD-14262, EMD-14263, EMD-14264, EMD-14265, and PDB-7R43 (Active-1092-iii); EMD- 14266, EMD14267, EMD-14268, EMD-14269, EMD-14270, and PDB-7R44 (Active-1092-iv); EMD-4272, EMD-14273, EMD-14274, EMD14275, EMD14276, and PDB-7R45 (Deactive-1092-i); EMD- 14277, EMD-14278, EMD14279, EMD-14280, EMD-14281, and PDB- 7R46 (Deactive-1092-ii); EMD-14282, EMD-14283, EMD-14284, EMD-14285, EMD-14286, and PDB-7R47 (Deactive-1092-iii); EMD- 14287, EMD-14288, EMD-14289, EMD-14290, EMD-14291, and PDB- 7R48 (Deactive-1092-iv); EMDB-14292, EMD-14293, EMD-14294, EMD-14295, EMD-14296, and PDB-7R4C (Deactive-1092-v); EMDB- 14297, EMD-14298, EMD-14299, EMD-14300, EMD-14301, and PDB- 7R4D (Deactive-1092-vi); EMDB-14302, EMD-14304, EMD-14305, EMD-14306, and PDB-7R4F (Slack-1092-i); EMD-14307, EMD- 14308, EMD-14309, EMD-14310, EMD-14311, and PDB-7R4G (Slack-1092-ii); EMBD-14127, EMD-14128, EMD-14129, EMD14130, EMD-14131, and PDB-7QSD (Inhibitor-free active); and EMDB-14126 (Inhibitor-free slack). Raw micrograph images are available at EMPIAR with accession codes EMPIAR-10991 (in presence of IM1761092) and EMPIAR-10984 (inhibitor-free). Otherwise, all data needed to evaluate the conclusions in the paper are present in the paper and/or the supplementary materials. IM1092 was supplied by ImmunoMet Therapeutics Inc. by collaboration agreement. ImmunoMet Therapeutics Inc. accepts proposals to supply IM1092 and ~1000 other biguanide compounds from their biguanide library for research purposes. Such proposals should be directed to dwelsch@immunomet.com. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/ science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.ade3332 Materials and Methods Supplementary Text Figs. S1 to S28 Tables S1 to S15 References (60–78) MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. Submitted 10 August 2022; accepted 23 December 2022 10.1126/science.ade3332 Bridges et al., Science 379, 351–357 (2023) 27 January 2023 7 of 7
10.1126_science.adf5848
RES EARCH R E S E A R C H A R T I C L E S U M M A R Y ◥ RENEWABLE ENERGY Declining cost of renewables and climate change curb the need for African hydropower expansion Angelo Carlino, Matthias Wildemeersch, Celray James Chawanda, Matteo Giuliani, Sebastian Sterl, Wim Thiery, Ann van Griensven, Andrea Castelletti* INTRODUCTION: Driven by population growth and the goal of improving living standards, especially in the least-developed regions, many African countries plan to expand their power generation capacities to meet future energy demand. Indeed, total electricity demand is expected to grow by 5 to 6% per year until 2050, mainly in sub-Saharan Africa. Yet the future of African energy systems will not only be driven by the additional energy demand but also by the need to mitigate and adapt to anthropogenic climate change. Hydropower is an important component of African power systems, especially in sub-Saharan countries. It provides around 20% of total electricity gen- eration, but its full potential has not been ex- ploited yet. Traditionally considered a cheap source of low-carbon electricity, more than 300 hydropower plants, corresponding to an additional 100-GW power capacity, are under consideration across the continent. RATIONALE: Although an apparently effective strategy, the long-term planning of hydropower systems is complex. First, as the cost of renew- ables continues to decline, solar and wind power are becoming more competitive and potentially cheaper alternatives. Second, in recent years, hydroclimatic variability has negatively affected hydropower generation in major river basins. Climate change will alter the spatiotemporal distribution of water availability, exacerbating the impacts of extreme events and reducing the predictability of future power generation. Finally, future energy demands and climate policies depend on evolving socioeconomic conditions that are fundamentally uncertain. In this work, we investigated the power capa- city expansion across the African continent over the next 30 years and elucidated the cost-optimal sequencing of hydropower projects. We built an integrated modeling framework that captures individual power project charac- teristics within an energy system model that simulates three socioeconomic scenarios that harmonize land-use change, climate impacts on water availability, final energy demands, and climate policy options. Our model relies on a combination of the Shared Socioeconomic Pathways (SSPs) and Representative Concen- Niger Nile Continent 10 0 ) W G ( y t i c a p a C ) W G ( y t i c a p a C 20 0 20 0 MED DRY MED DRY Congo Zambezi 10 0 MED DRY MED DRY 100 ) W G ( y t i c a p a C 50 0 MED DRY River basins Proposed hydropower projects Sustainability Inequality Fossil-fuel development Cost-optimal hydropower expansion. Proposed (dashed line) and cost-optimal (bars) capacity expansion for continental Africa and its major river basins under the scenarios considered. In total, 32 to 60% of the proposed capacity is not cost-optimal. More than half of the capacity proposed for the Nile, Congo, and Niger basins is always cost-optimal, whereas the expansion in the Zambezi River basin depends on the considered scenario. The colors of the shaded areas in the map correspond to the river basins represented by each graph. tration Pathways, namely SSP1-2.6, SSP4-6.0, and SSP5-8.5. SSP1-2.6 describes a sustain- able development scenario that aims to main- tain the global mean temperature below 2°C, whereas the other two scenarios result in higher levels of warming and are characterized by rising inequalities and fossil-fueled develop- ment, respectively. We considered median (MED) and very dry (DRY) water availability scenarios to capture hydroclimatic variability reflecting a risk-neutral and risk-averse plan- ning perspective. RESULTS: Our results show that between 32 and 60% of the proposed hydropower capacity is not cost-optimal. Moreover, our analysis suggests that hardly any new hydropower will be built after 2030, meaning that its role in terms of installed capacity and generation will gradually decrease in favor of solar and wind power. Across the scenarios, hydropower expan- sion is robust in the Nile, Congo, and Niger river basins, whereas it remains uncertain in the Zambezi and smaller river basins. These find- ings emphasize the importance of connecting hydropower planning with capacity expansion models, because cost-optimality cannot be deter- mined solely based on each project’s technical characteristics. Finally, we discover that an in- crease in annual capital investment between 1 and 4% at the continental level can ensure the reliability of the power system against hydro- climatic variability. Yet the required increase in capital investments and the observed reduc- tions in vulnerability do not necessarily overlap at the country level. As local interests conflict and diverge from system-wide ones, we under- line the importance of electricity exchanges between countries and cooperation for power system reliability. CONCLUSION: Traditional planning of hydro- power facilities is challenged by the dynamics of technological innovation, climate impacts on water availability, and uncertainty in long-term socioeconomic projections that affect energy demands and climate policies. Using a multi- sectoral modeling framework, we designed capa- city expansion plans that avoid commitments to cost-inefficient hydropower infrastructures that are often associated with substantial impacts on the local communities and environment. Yet, in the short term, especially in the transition to a net-zero emissions energy system, hydropower represents a cheap alternative to displace fossil fuels, especially coal.▪ The list of author affiliations is available in the full article online. *Corresponding author. Email: andrea.castelletti@polimi.it Cite this article as A. Carlino et al., Science 381, eadf5848 (2023). DOI: 10.1126/science.adf5848 READ THE FULL ARTICLE AT https://doi.org/10.1126/science.adf5848 Carlino et al., Science 381, 645 (2023) 11 August 2023 1 of 1 RES EARCH R E S E A R C H A R T I C L E ◥ RENEWABLE ENERGY Declining cost of renewables and climate change curb the need for African hydropower expansion Angelo Carlino1,2†, Matthias Wildemeersch2‡, Celray James Chawanda3, Matteo Giuliani1, Sebastian Sterl3,4, Wim Thiery3, Ann van Griensven3, Andrea Castelletti1* Across continental Africa, more than 300 new hydropower projects are under consideration to meet the future energy demand that is expected based on the growing population and increasing energy access. Yet large uncertainties associated with hydroclimatic and socioeconomic changes challenge hydropower planning. In this work, we show that only 40 to 68% of the candidate hydropower capacity in Africa is economically attractive. By analyzing the African energy systems’ development from 2020 to 2050 for different scenarios of energy demand, land-use change, and climate impacts on water availability, we find that wind and solar outcompete hydropower by 2030. An additional 1.8 to 4% increase in annual continental investment ensures reliability against future hydroclimatic variability. However, cooperation between countries is needed to overcome the divergent spatial distribution of investment costs and potential energy deficits. O ver the next few decades, African energy systems are expected to undergo pro- found changes. The total electricity de- mand is predicted to increase by 5 to 6% per year over the next 10 years until 2050 (1–3), an increase that is driven by sustained population growth, mainly in sub-Saharan Africa (4), and the continuous infrastructural investments aimed at improving energy access and living standards, especially in the least- developed areas (5, 6). This increasing demand, together with the need to mitigate and adapt to anthropogenic climate change (7), will shape the future development of African energy sys- tems. The use of low-carbon energy sources (3, 8, 9) will gradually lessen the historical dependency on fossil fuels, which are abun- dant across the continent (10). In the short term, annual investments of US$190 billion are required to ensure such a successful energy transition, with more than two-thirds of this financial investment allocated to clean-energy sources (3). Among these, hydropower has historically been favored as a low-cost source of baseload power (11), and policies that are in place now imply a substantial infrastructural expansion (12). Moreover, hydropower is an attractive component of the future African power system owing to its ability to balance 1Department of Electronics, Information, and Bioengineering, Politecnico di Milano, Milano, Italy. 2International Institute for Applied Systems Analysis, Laxenburg, Vienna, Austria. 3Department of Hydrology and Hydraulic Engineering, Vrije Universiteit Brussel, Brussels, Belgium. 4World Resources Institute, Regional Hub for Africa, Addis Ababa, Ethiopia. *Corresponding author. Email: andrea.castelletti@polimi.it †Present address: Department of Global Ecology, Carnegie Institution for Science, Stanford, CA, USA. ‡Present address: Environmental Change Institute, Oxford University, Oxford, UK. grid load in support of intermittent renewable electricity sources (13–15) and because the remaining untapped potential across the con- tinent is relatively large (11). According to plans of national and regional agencies, more than 300 new hydropower projects are, at present, committed, planned, or under consid- eration across the African continent (16). These projects amount to a total of around 100 GW of additional hydropower capacity, with 168 large (≥100-MW) projects accounting for almost 90 GW (16). Nevertheless, climate change makes future hydropower generation uncertain (17) and in- creases the risk of cascading power system failures across countries and power pools (18), likely jeopardizing its potential to foster resil- ience (19). Moreover, capacity expansion pro- jections are linked to future energy demand, technology costs, and climate policy, which are fundamentally uncertain factors (20, 21). The excessive reliance on hydropower in many sub- Saharan countries is presently a source of con- cern and a reason for caution in additional hydropower investment (22). Further doubts are cast on hydropower capacity expansion (23) when socioeconomic and environmental impacts of hydropower are analyzed, such as population displacement (24), reduced sediment connectivity (25), loss of biodiversity (26), and competition with other water uses, most impor- tantly with agriculture (21). Given the scale of future infrastructure devel- opment, the socioeconomic and environmental impacts of hydropower expansion, and the need to bridge continental as well as regional power system development, it is crucial to identify the hydropower projects that should be prioritized and the ones that should be discarded based on the cost-optimal power system capacity expansion. Indeed, the selection and sequencing of the required hydropower infrastructure, given energy, socioeconomic, and technological development, is a critical first step. Further research should evaluate the ensuing social, climatic, and environmental impacts on the alternatives of interest to sup- port final planning decisions. To what extent do the planned hydropower expansion and its spatial distribution over the main river basins change depending on socioeconomic, land-use, and climatic uncertainties? What are the costs of climate-proofing the energy system, and how are these costs spatially distributed compared with power deficits driven by hydroclimatic variability? In this work, we build an integrated model- ing framework to examine the role of hydro- power in a sustainable energy transition that is cognizant of hydroclimatic and land-use change, socioeconomic projections, and climate policy options. Although previous studies on strategic dam planning (27–30) rarely included the power system and rarely went beyond the basin scale (31, 32), our analysis examines the full energy portfolio at the continental scale. Specifically, we consider cross-basin interac- tions across the power grid (33), hydropower projects proposed at the river basin and national scales, and socioeconomic and land-use projec- tions. By doing so, we limit undesirable out- comes that result from the integration of national, regional, and continental policies across multi- ple sectors and scales (34). Our results show that hydropower will have lost its dominant role in Africa’s renewable electricity mix by 2050, with solar and wind power representing at least 29 to 38% and 8 to 12% of generation, respectively, and hydro- power’s share shrinking to 7 to 14% under all considered scenarios. Between 40 and 68% of the proposed new hydropower capacity or, in other words, between 120 and 251 of the 367 proposed projects could potentially be cost- optimal, and nearly no new hydropower plants are recommended to be built after 2030. Al- though the viability of hydropower expansion in the Zambezi River basin is dependent on the scenario that is considered, many of the proposed projects for the Nile, Congo, and Niger remain economically viable under all considered scenarios. Finally, guaranteeing the reliability of the energy system against hydroclimatic risks only requires reallocat- ing some of the investments in hydropower toward other sources, especially solar power and firming technologies, with a small in- crease in annual capital investments. Yet the need for additional investment and the risk of shortages are often located in different regions. As a consequence, we highlight the importance of transnational governance mea- sures to guarantee climate-resilient energy systems. Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E Sequencing hydropower projects within power capacity expansion To obtain plans for hydropower project se- quencing and associated power capacity ex- pansion, we set up a multiscale, multisector modeling approach (fig. S1). We combined input data from three main datasets. First, we used the Shared Socioeconomic Pathways (SSPs) database (35) to obtain projected energy demands. Second, we relied on the African Hydropower Atlas (16) to characterize each hydropower project in the OSeMOSYS-TEMBA model (36). Third, to coherently account for the coevolution of the climatic and the socio- economic system, we used the Inter-Sectoral Impact Model Intercomparison Project (ISI- MIP2b) scenarios (37) to represent the future hydrological regime that will result from changes in the climate system and the land-use sector. Natural climate variability was considered using a median and a very dry hydrological scenario. These correspond to the 50th and 5th percen- tiles of the distribution of simulated annual average generation, which is obtained by simu- lating a distributed hydrological model under an ensemble of climate projections from 2020 to 2050 (see materials and methods). We used this model to study the expansion trajectory of the African energy systems over the period from 2020 to 2050 at the continen- tal scale assuming centralized decision-making. We considered three socioeconomic scenarios that aggregate socioeconomic, land-use, and climatic assumptions: (i) a sustainable develop- ment scenario, using a carbon emission con- straint compatible with a 2°C long-term warming, according to SSP1-2.6; (ii) a scenario designed to focus on heterogeneous economic develop- ment among regions not associated with climate policy efforts, according to SSP4-6.0; and (iii) a fossil-fueled economic growth scenario asso- ciated with high greenhouse gas emissions, according to SSP5-8.5. For each socioeconomic scenario, we consider the median (MED) and very dry (DRY) hydrological scenarios. We use the first to represent traditional hydropower planning and the second to stress-test the power system under worst-case hydroclimatic conditions. Indeed, these two scenarios can be seen as describing different risk-preparedness targets (risk–neutral and risk-averse, respectively) with respect to the uncertainty associated with hydroclimatic variability. For each considered scenario, we optimized the power capacity ex- pansion for each energy source and the sequenc- ing of the proposed (i.e., planned, committed, and candidate) hydropower projects collected in the African Hydropower Atlas. Moreover, we examined the cost-reliability trade-off at differ- ent spatial scales, which would otherwise remain hidden behind the large-scale formulation of the least-cost capacity expansion problem. Cost-effectiveness of solar energy avoids the need for long-term hydropower expansion Our model results show that at least one-third of the new hydropower capacity proposed at the regional and country levels is not cost- optimal across continental Africa, and this result holds under all considered scenarios (Fig. 1). Under ensemble median hydrologic change (i.e., under the MED scenarios), new hydropower installed capacity ranges from 52 GW under SSP4-6.0 to 66 GW under SSP1-2.6, whereas these values drop to between 39 GW (SSP4-6.0) 47 GW (SSP1-2.6) when considering dry hydrology conditions under the risk-averse approach (i.e., under the DRY scenarios), meaning that more than half of the proposed capacity is not economically viable at the continental scale. In all these plans, two large projects are responsible for more than 17 GW of viable capacity: the soon to be completed 6.4-GW Grand Ethiopian Renaissance Dam and the 11.0-GW Inga 3 candidate project in the Democratic Republic of the Congo. In general, the SSP1-2.6 scenario consistent with a warming of 2°C at the global level requires more hydro- power than other scenarios owing to the reduced reliance on fossil fuels. To isolate the impact of climate change on hydropower expansion, we examined capacity expansion strategies by considering hydropower generation based on observations from 1986 to 2005. We see that climate change is particularly affecting the scenarios with the largest hydropower expan- sion and is responsible for a reduction of 9 GW (SSP1-2.6) and 8 GW (SSP5-8.5) (fig. S2). As we consider the salvage value of infrastructure at the end of the planning horizon that corresponds with the remaining operational life, our results remain consistent when we extend the horizon until 2070 (fig. S3). Under all socioeconomic and hydrological scenarios, at least half of the additional hydro- power capacity is installed in the period from 2020 to 2030 (Fig. 1, A to C), with the window in which hydropower can still compete econo- mically with solar photovoltaics (PV) rapidly closing. Beyond 2030, the share of new invest- ments in solar power increases substantially, and further development of hydropower in Africa is unlikely to be cost-effective (Fig. 2). Although hydropower could still be competitive with solar PV until the end of this decade, the often-witnessed build time and cost overruns for hydropower projects (38) may even pre- clude large-scale hydropower expansion before that time, paving the way for further solar PV deployment. In addition, all capacity invest- ments are expected to grow rapidly in the decades after 2030, thus further diminishing the role of hydropower in the future energy portfolio (39). Similarly, given the large expan- sion of the power system in the next few decades, the decline of hydropower is also substantial in the total capacity share (fig. S4). The gap is even more pronounced for the DRY scenarios in which more than half of the Fig. 1. Decadal and total hydropower capacity expansion under the considered scenarios. The dashed line indicates the capacity of proposed projects reported in the African Hydropower Atlas. The colors of the bars are associated with the considered SSP scenarios, which coherently capture socioeconomic, land-use, and hydroclimatic change. For each decade and for the total, the bars on the left report the capacity expansion plan designed under ensemble median hydrology (MED), and the bars on the right correspond to the capacity expansion plan designed under dry hydrology (DRY). The arrows indicate the fraction of proposed capacity, which is not cost-optimal under the two different risk-preparedness targets. Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Power capacity expansion at the continental level. The share of new installed capacity for each power source (left y axis) and the new installed capacity (right y axis) are reported against the three decades examined. The top row reports the share of each power source in new capacity under median hydrology (MED scenarios), and the bottom row reports the capacity expansion plans designed under dry hydrology (DRY scenarios). Each column represents the different SSP scenarios. Each power source is described in the legend, the order of which, from top to bottom, follows the order of the stacked bars from top to bottom. CCS, carbon capture and storage; CSP, concentrating solar power. proposed capacity is not economically optimal, resulting in higher investments in solar pow- er (bottom row in Fig. 2). Solar power becomes a option, displacing hydropower projects whose generation changes the most from median to very dry hydrology. In SSP1-2.6, an important role is played by nuclear power by mid-century, which is used to further reduce investment in fossil fuel power sources and represents an important share of energy generation in 2050 (fig. S5). Contrary to SSP1-2.6, where coal be- comes almost absent, under SSP4-6.0 and SSP5-8.5, it still contributes around 40% of the energy generation mix by mid-century, with more than 2000 terawatt-hours under SSP5-8.5 (fig. S5). With respect to the flexibility required in the power system to balance the reduced output of solar plants at night, hydro- power comes after biomass and fossil fuels, and wind has a complementary diurnal profile to solar as well (fig. S6). Consequently, our results do not suggest that hydropower will still be a major provider of firm generation and flexibility by mid-century. Location and drivers of hydropower expansion Most of the planned African hydropower pro- jects concentrate in four major river basins— Nile, Congo, Zambezi, and Niger—which account for around 66% of the total proposed additio- nal hydropower capacity (16). Across the socio- economic and risk-preparedness scenarios, the cost-optimal dam portfolio varies substan- tially, even though some projects are consistently selected (fig. S7). A robust finding over the considered scenarios and river basins is that less hydropower is installed in the DRY capa- city expansion scenarios and under SSP4-6.0 and SSP5-8.5 (Fig. 3). The Congo River basin is consistently cost-optimal for around half of its potential through the Inga 3 Dam, accounting for 11 GW in the Democratic Republic of the Congo and built in all the scenarios. Half of the proposed potential for the Nile River basin is always cost-optimal, mainly in Ethiopia and Uganda, up to 80% in SSP5-8.5 with MED capacity expansion. The hydropower expan- sion in the Zambezi River basin is instead very uncertain and strongly dependent upon the considered scenario, ranging from 30% (SSP5- 8.5) to 70% (SSP1-2.6) of the proposed capacity in the MED scenarios and between 13% (SSP4- 6.0 and SSP5-8.5) and 39% (SSP1-2.6) in the DRY scenarios. Finally, the cost-optimal hydro- power potential in the Niger River basin is between 86% (SSP4-6.0) and 91% (SSP1-2.6) of the proposed capacity for the MED scenarios, and it is reduced to between 53% (SSP4-6.0) and 83% (SSP5-8.5) in the DRY scenarios. These projects are located mainly in Nigeria, a po- tential hotspot of hydropower development. For what concerns the remaining smaller basins, the development of projects varies considerably from 38% (SSP4-6.0) to 71% (SSP1-2.6) of their total capacity in the MED capacity expansion scenarios and between 24% (SSP4-6.0) and 32% (SSP1-2.6) for the DRY capacity expan- sion scenarios. Given these results, we can partially trace the cost-optimal power expansion decisions back to the characteristics of the proposed hydro- power projects. High average capacity factors and high capacity are usually good indicators of cost-optimality (Fig. 4). Indeed, the higher the capacity, the lower the capital cost of new hydropower (40), even though the probabilities of delays and cost overruns increase as well (41). Furthermore, the higher the average capacity factor, the higher the annual generation of a power plant. The construction of new hydro- power projects is not sensitive to the interannual variability in the capacity factor. Then again, spatial and temporal energy system constraints, such as transmission line capacity and prox- imity to more economically favorable hydro- power projects, enable a full understanding of the cost-optimal power system development. This is, for example, the case for projects in the Zambezi basin in Zambia, a region well con- nected to the Democratic Republic of the Congo. The development of the Inga 3 Dam in the latter allows for substantial cheap electricity exports to neighboring countries, reducing the viability of domestic hydropower expansion in Zambia. The regional distribution of costs and deficits requires cooperation It is presently unclear how the magnitude of drought-induced power deficits compares with the size of additional investment costs that are required to climate-proof the energy system Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E SSP1-2.6 SSP4-6.0 SSP5-8.5 UG ET 15.3 23.1 22.0 Full Potential 11.5 31.0 10.0 2 2.1 CM AO 9.1 NG MED 1 2.2 CD DRY 1 3.0 CD 7.7 Z M ET 14.0 23.1 31.0 10.0 Full Potential 22.0 11.5 5 . 4 Z M UG ET 13.7 23.1 1 2.1 CD 22.0 Full Potential 31.0 10.0 11.5 1 . 5 M Z 8.6 NG ET 18.5 23.1 UG 31.0 10.0 22.0 Full Potential 11.5 1 6 . 0 ET 8.9 NG 3 . 2 1 1 2.1 CD 3 . 3 ET 14.6 23.1 ET 13.1 23.1 22.0 31.0 10.0 Full Potential 11.5 4 . 0 1 1 1.7 CD 5.4 NG 5 . 1 7.5 1 1.7 CD 22.0 5.3 NG 5 . 1 31.0 10.0 Full Potential 11.5 8.0 8.3 NG Fig. 3. Basin- and country-level hydropower capacity expansion. The full capacity of proposed projects is reported by river basin in the inner ring. The cost-optimal capacity in each river basin is reported in the middle ring and is assigned to the corresponding countries in the outer ring. All values are reported in GW units. The columns correspond to the SSP scenarios examined. The top and bottom rows report results from the MED and DRY capacity expansion scenarios, respectively. For each scenario, only countries that are building more than 2.5 GW of new hydropower are labeled, namely Angola (AO), Democratic Republic of the Congo (CD), Cameroon (CM), Ethiopia (ET), Mozambique (MZ), Nigeria (NG), Uganda (UG), and Zambia (ZM). (i.e., to guarantee demand satisfaction under dry hydrological conditions). For this reason, we stress-test the MED capacity expansion plans, obtained under median hydrology, by simulat- ing it under dry hydrology to estimate the po- tential deficit that can occur. The observed generation deficits should be understood as the result of planning the power capacity expan- sions for each source, not only for hydropower, without explicitly accounting for hydroclimatic variability. The reported deficits present a worst- case scenario because safety mechanisms such as reserve margins are supposed to be in place to reduce the probability of occurrence and the magnitude of these events. The DRY capacity expansion plans can remove this risk with a capital cost increase between 1.8% (SSP5-8.5) and 4% (SSP1-2.6) in annual capital investments at the continental level under all the socio- economic scenarios. Yet at the country level, the cost increase and potential deficit are un- evenly distributed and vary widely across the scenarios (Fig. 5). Generally, reduced hydropower generation requires backing up with existing, mainly fossil fuel–based technologies or with additional capa- city. This additional capacity is typically solar PV under cost-optimal expansion scenarios, especially under SSP1-2.6, in which the reliance on fossil fuels for power generation is con- strained. Consequently, spatial planning of the deployment of renewable power plants will be affected as well. For many regions that are not dependent on hydropower, there is no difference between the two plans because they are not affected by power deficits or additional costs induced by hydrological variability (Northern Africa and South Africa). Nonetheless, power pools strongly dependent on hydropower, such as the Southern African, Eastern Africa, and West African Power Pools, are more subject to cost increase and power deficit. Under SSP1-2.6, West Africa is affected by generation deficit events that require substantial capital investments to en- sure reliability (e.g., Senegal, Guinea-Bissau, Ghana, and Togo). Conversely, the power de- ficits in Nigeria and Burkina Faso require a modest increase in annual capital cost. In the other scenarios, the power deficit affects most- ly Mali, Niger, and Benin, but the costs to achieve reliability remain low in all the power pool. With respect to the Eastern Africa Power Pool, Ethiopia, Tanzania, Uganda, Rwanda, and South Sudan are most at risk of power outages induced by hydroclimatic variability. All of these coun- tries require substantial investments to reduce this risk, and additional economic efforts will be required from Egypt, Sudan, and Kenya, especially in the case of SSP1-2.6. With respect to the Southern African Power Pool and scenario SSP1-2.6, Zambia, Namibia, and Mozambique remain most vulnerable to droughts. Zambia is particularly at risk because the power deficit would be around 13%, which could be miti- gated with an 11% increase in annual capital in- vestment. In addition to the above-mentioned countries, in the Central African Power Pool, Angola, Zimbabwe, and the neighboring Demo- cratic Republic of the Congo are also required to increase their investments to climate-proof their energy systems to a substantial extent. Under the other scenarios, Zambia always remains ex- posed to drought-related power outage risk, together with Namibia, whose cost to ensure reliability remains lower. In all scenarios, a generation deficit is observed if power trade is not allowed between countries, underscoring the importance of cooperation and political stability in the region (fig. S8). Discussion As African power demand grows, especially in sub-Saharan Africa, the remaining untapped hydropower potential represents a cheap, clean energy source, which explains the large number of infrastructural projects that are presently under consideration. However, as costs associ- ated with solar and wind power generation continue to decline, the historical reliance on hydropower of many sub-Saharan African coun- tries might come to an end. Solar and wind power are expected to become the primary power sources in 2050, representing 50% of the electricity mix of the continent in the sustaina- ble development scenario compatible with a 2°C long-term warming (SSP1-2.6) and always representing at least 50% of new installed capacity in the next three decades under all scenarios considered. Even under the SSP1-2.6 scenario, which pushes for extensive renewable capacity expansion, no more than 67% of pro- posed hydropower capacity is cost-optimal, and this percentage shrinks to 48% under the assumption of aversion to hydroclimatic risk. Project delays and cost overruns might further favor solar and wind projects, making hydro- power development even less competitive from an economic perspective (42). Yet in the short term, especially in the transition to a final net- zero configuration, hydropower represents a cheap alternative to avoid the high costs of installing solar and wind at the present level of technological maturity and to displace fossil fuels, mainly coal. The Nile, Congo, and Niger River basins provide reliable hydropower gener- ation. Yet the development of projects in these regions needs to be accompanied by investment in grid capacity in order to reap all the benefits of large hydropower. Climate-proofing the energy system against hydroclimatic variability requires reducing investment in hydropower Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Main characteristics of projects and their role in least-cost capacity expan- sion. Capacity, average capacity factor under very dry hydrology, and maximum interannual capacity factor variability under very dry hydrology of the examined hydropower projects are reported on the x axis, the y axis, and with different colors, respectively. The diamond marker indicates a project that is always cost-optimal. The circles correspond to projects that are cost-optimal at least once, whereas the crosses correspond to the projects that are never built. The always– cost-optimal projects correlate with the average capacity factor (point biserial correlation coefficient = 0.37; p = 2 × 10−19) and capacity (point biserial correlation coefficient = 0.18; p = 3 × 10−5). Their correlation with capacity factor's variability is weaker (point biserial correlation coefficient = 0.09; p = 0.03). Fig. 5. Country-level cost-deficit trade-offs. Maximum annual power deficit as a percentage of demand over the period from 2020 to 2050 obtained from simulation of the MED capacity expansion plan under dry hydrology. The additional cost of eliminating the power deficits is derived as the percentage increase obtained from the annualized capital costs of the MED and the DRY capacity expansion plans. Their joint value is reported for each country in the maps by the bidimensional color scale that is visible in the legend, and each map corresponds to a different SSP scenario. and investing in additional solar, wind, and firm- ing capacity, particularly in the scenarios where emissions are constrained. These additional costs are not necessarily distributed uniformly or fairly across the countries, highlighting the need for coordination and incentive mecha- nisms to support capacity expansion plans, which are robust to climate change impacts. Through the reduction in economically viable hydropower capacity associated with the declining cost of wind and solar, techno- logical innovation helps reduce pressure on riverine ecosystems and small communities in the proximity of proposed impoundments and further downstream as far as the impacts of these changes propagate (43). Indeed, previous research on hydropower’s social and environ- mental trade-offs (25, 27, 30) and the effects of environmental risks on the financial per- formance of this infrastructure (44) has sug- gested caution in construction of new projects. Introducing these factors into our modeling framework is likely to further reduce the space for hydropower in future energy systems. Analy- ses at the river-basin level remain complemen- tary to our study and might be better tailored to address such concerns. However, additional research and development of new methods are needed to connect local, regional, and continen- tal scales for a robust planning of water and energy systems (34). Similarly, greenhouse gas emissions from reservoirs (30, 45–47) are a de- terrent for hydropower capacity expansion, particularly in tropical areas where life-cycle emissions associated with new dams might be comparable to those of fossil fuel power sources (48, 49). Accounting for this factor will likely further promote the expansion of wind, solar, and other carbon-neutral technologies. Additionally, we are not able to fully capture the contribution of hydropower projects to ancillary services such as frequency regulation and improved renewable integration associ- ated with the rapid ramp-up of power output. Although these services are rarely considered Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E in hydropower planning, their importance will rise as more wind power and solar power are added to the grid, potentially affecting our results. Moreover, electricity generation is not always the main purpose for which water res- ervoirs are built. If some of the reservoir hydro- power projects were to be associated with other needs (e.g., agriculture, flood control, drinking water supply), cross-sectoral interactions could improve their economic performance and make them attractive investments. In this case, reser- voir greenhouse gas emissions should not be at- tributed to electricity generation only, but the exact attribution of greenhouse gas emissions to the different sectors remains a complex issue. Governance and political stability are key to ensuring sustainable exploitation of the eco- nomically viable hydropower potential, parti- cularly in transboundary river basins (50). The Nile and the Niger River basins, identified as hotspots of hydropower development, are high- risk areas because of their transboundary nature in regions of political instability and presence of armed conflict (51). Implementation of coop- eration schemes is crucial to reduce tensions and provide water and energy security in these areas (13, 15, 52–54). In a broader sense, cooperation and govern- ance are fundamental to allow all African countries to switch their focus from energy independence to energy security (55). In this regard, establishing power pools and the Africa Clean Energy Corridor has been crucial for en- ergy governance. These mechanisms and in- vestments paved the way for increased energy security across the continent (56). To prepare for the impacts of dry years, investment in alternative power sources is required, even in locations that might not be directly affected by generation deficits. Understanding the con- sequences of interconnected power systems can therefore promote the design of agree- ments and policy interventions that foster energy security and resilience in the face of hydroclimatic change. Growing evidence moti- vates concerns about the increased risk of conflict and instability associated with the growing impacts of climate change (57). Govern- ments and power pools must prepare for stressful contexts where local strategies do not match large-scale cost-optimal develop- ment. To confront the friction between coor- dinated and decentralized decision-making levels, mechanisms building on incentive schemes and side payments need to be designed. In this conundrum, our results can inform future research to ensure multiscale coordination for energy security and sustainable hydropower development within the African continent. RE FE RENCES AND N OT ES 1. N. S. Ouedraogo, Africa energy future: Alternative scenarios and their implications for sustainable development strategies. Energy Policy 106, 457–471 (2017). doi: 10.1016/ j.enpol.2017.03.021 2. 3. I. Pappis, M. Howells, V. Sridharan, W. Usher, A. Shivakumar, F. Gardumi, E. Ramos, “Energy projections for African countries,” G. Hidalgo, H. Medarac, R. González-Sánchez, I. Kougias, Eds. (JRC Technical Report JRC118432, Publications Office of the European Union, 2019). doi: 10.2760/678700 International Energy Agency (IEA), “Africa energy outlook 2022” (IEA, 2022); https://iea.blob.core.windows.net/assets/ 220b2862-33a6-47bd-81e9-00e586f4d384/ AfricaEnergyOutlook2022.pdf. 4. United Nations Department of Economic and Social Affairs, “World population prospects 2022: Summary of results” (United Nations, 2022); https://www.un.org/development/ desa/pd/sites/www.un.org.development.desa.pd/files/ wpp2022_summary_of_results.pdf. 5. N. D. Rao, J. Min, A. Mastrucci, Energy requirements for decent living in India, Brazil and South Africa. Nat. Energy 4, 1025–1032 (2019). doi: 10.1038/s41560-019-0497-9 6. F. Dalla Longa, B. van der Zwaan, Heart of light: An assessment of enhanced electricity access in Africa. Renew. Sustain. Energy Rev. 136, 110399 (2021). doi: 10.1016/ j.rser.2020.110399 7. C. Trisos, I. O. Adelekan, E. Totin, A. Ayanlade, J. Efitre, A. Gemeda, K. Kalaba, C. Lennard, C. Masao, Y. Mgaya, G. Ngaruiya, D. Olago, N. P. Simpson, S. Zakieldeen, “Africa” in Climate Change 2022: Impacts, Adaptation and Vulnerability, H.-O. Pörtner, D. C. Roberts, M. M. B. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegria, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, B. Rama, Eds. (Cambridge Univ. Press, 2022), pp. 1285–1455. 8. B. van der Zwaan, T. Kober, F. Dalla Longa, A. van der Laan, G. J. Kramer, An integrated assessment of pathways for low-carbon development in Africa. Energy Policy 117, 387–395 (2018). doi: 10.1016/j.enpol.2018.03.017 9. D. Puig et al., An action agenda for Africa’s electricity sector. Science 373, 616–619 (2021). doi: 10.1126/science.abh1975; pmid: 34353936 10. J. C. Steckel, J. Hilaire, M. Jakob, O. Edenhofer, Coal and carbonization in sub-Saharan Africa. Nat. Clim. Change 10, 83–88 (2020). doi: 10.1038/s41558-019-0649-8 11. D. E. Gernaat, P. Bogaart, D. P. V. Vuuren, H. Biemans, R. Niessink, High-resolution assessment of global technical and economic hydropower potential. Nat. Energy 2, 821–828 (2017). doi: 10.1038/s41560-017-0006-y 12. C. Zarfl, A. Lumsdon, J. Berlekamp, L. Tydecks, K. Tockner, A global boom in hydropower dam construction. Aquat. Sci. 77, 161–170 (2015). doi: 10.1007/s00027-014-0377-0 13. S. Sterl et al., Smart renewable electricity portfolios in West Africa. Nat. Sustain. 3, 710–719 (2020). doi: 10.1038/ s41893-020-0539-0 14. E. G. Dimanchev, J. L. Hodge, J. E. Parsons, The role of hydropower reservoirs in deep decarbonization policy. Energy Policy 155, 112369 (2021). doi: 10.1016/j.enpol.2021.112369 15. S. Sterl, D. Fadly, S. Liersch, H. Koch, W. Thiery, Linking solar and wind power in eastern Africa with operation of the Grand Ethiopian Renaissance Dam. Nat. Energy 6, 407–418 (2021). doi: 10.1038/s41560-021-00799-5 16. S. Sterl et al., A spatiotemporal atlas of hydropower in Africa for energy modelling purposes. Open Res. Eur. 1, 29 (2022). doi: 10.12688/openreseurope.13392.3 17. S. G. Yalew et al., Impacts of climate change on energy systems in global and regional scenarios. Nat. Energy 5, 794–802 (2020). doi: 10.1038/s41560-020-0664-z 18. D. Conway, C. Dalin, W. A. Landman, T. Osborn, Hydropower plans in eastern and southern Africa increase risk of concurrent climate-related electricity supply disruption. Nat. Energy 2, 946–953 (2017). doi: 10.1038/s41560-017-0037-4 19. V. Sridharan et al., Resilience of the Eastern African electricity sector to climate driven changes in hydropower generation. Nat. Commun. 10, 302 (2019). doi: 10.1038/s41467-018- 08275-7; pmid: 30655521 20. P. A. Ray et al., Multidimensional stress test for hydropower investments facing climate, geophysical and financial uncertainty. Glob. Environ. Change 48, 168–181 (2018). doi: 10.1016/j.gloenvcha.2017.11.013 21. X. Zhang et al., Impacts of climate change, policy and water- energy-food nexus on hydropower development. Renew. Energy 116, 827–834 (2018). doi: 10.1016/j.renene.2017.10.030 22. G. Falchetta, D. E. Gernaat, J. Hunt, S. Sterl, Hydropower dependency and climate change in sub-Saharan Africa: A nexus framework and evidence-based review. J. Clean. Prod. 231, 1399–1417 (2019). doi: 10.1016/j.jclepro.2019.05.263 23. E. F. Moran, M. C. Lopez, N. Moore, N. Müller, D. W. Hyndman, Sustainable hydropower in the 21st century. Proc. Natl. Acad. Sci. U.S.A. 115, 11891–11898 (2018). doi: 10.1073/ pnas.1809426115; pmid: 30397145 24. A. Mayer, L. Castro-Diaz, M. C. Lopez, G. Leturcq, E. F. Moran, Is hydropower worth it? Exploring amazonian resettlement, human development and environmental costs with the Belo Monte project in Brazil. Energy Res. Soc. Sci. 78, 102129 (2021). doi: 10.1016/j.erss.2021.102129 25. R. J. P. Schmitt, S. Bizzi, A. Castelletti, J. J. Opperman, G. M. Kondolf, Planning dam portfolios for low sediment trapping shows limits for sustainable hydropower in the Mekong. Sci. Adv. 5, eaaw2175 (2019). doi: 10.1126/sciadv. aaw2175; pmid: 32047852 26. V. Barbarossa et al., Impacts of current and future large dams on the geographic range connectivity of freshwater fish worldwide. Proc. Natl. Acad. Sci. U.S.A. 117, 3648–3655 (2020). doi: 10.1073/pnas.1912776117; pmid: 32015125 27. R. J. Schmitt, S. Bizzi, A. Castelletti, G. M. Kondolf, Improved trade-offs of hydropower and sand connectivity by strategic dam planning in the Mekong. Nat. Sustain. 1, 96–104 (2018). doi: 10.1038/s41893-018-0022-3 28. A. Vinca, K. Riahi, A. Rowe, N. Djilali, Climate-land-energy-water nexus models across scales: Progress, gaps and best accessibility practices. Front. Environ. Sci. 9, 691523 (2021). doi: 10.3389/fenvs.2021.691523 29. R. M. Almeida et al., Strategic planning of hydropower development: Balancing benefits and socioenvironmental costs. Curr. Opin. Environ. Sustain. 56, 101175 (2022). doi: 10.1016/j.cosust.2022.101175 30. A. S. Flecker et al., Reducing adverse impacts of Amazon hydropower expansion. Science 375, 753–760 (2022). doi: 10.1126/science.abj4017; pmid: 35175810 31. R. J. Schmitt, N. Kittner, G. M. Kondolf, D. M. Kammen, Joint strategic energy and river basin planning to reduce dam impacts on rivers in Myanmar. Environ. Res. Lett. 16, 054054 (2021). doi: 10.1088/1748-9326/abe329 32. K. Siala, A. K. Chowdhury, T. D. Dang, S. Galelli, Solar energy and regional coordination as a feasible alternative to large hydropower in Southeast Asia. Nat. Commun. 12, 4159 (2021). doi: 10.1038/s41467-021-24437-6; pmid: 34230491 33. A. L. Cáceres, P. Jaramillo, H. S. Matthews, C. Samaras, B. Nijssen, B., Potential hydropower contribution to mitigate climate risk and build resilience in Africa. Nat. Clim. Change 12, 719–727 (2022). doi: 10.1038/s41558-022-01413-6 34. M. Giuliani, J. R. Lamontagne, M. I. Hejazi, P. M. Reed, A. Castelletti, Unintended consequences of climate change mitigation for African river basins. Nat. Clim. Change 12, 187–192 (2022). doi: 10.1038/s41558-021-01262-9 35. K. Riahi et al., The Shared Socioeconomic Pathways and their energy, land use, and greenhouse gas emissions implications: An overview. Glob. Environ. Change 42, 153–168 (2017). doi: 10.1016/j.gloenvcha.2016.05.009 36. I. Pappis et al., The effects of climate change mitigation strategies on the energy system of Africa and its associated water footprint. Environ. Res. Lett. 17, 044048 (2022). doi: 10.1088/1748-9326/ac5ede 37. K. Frieler et al., Assessing the impacts of 1.5 °C global warming–simulation protocol of the Inter-Sectoral Impact Model Intercomparison Project (ISIMIP2b). Geosci. Model Dev. 10, 4321–4345 (2017). doi: 10.5194/gmd-10-4321-2017 38. B. K. Sovacool, A. Gilbert, D. Nugent, An international comparative assessment of construction cost overruns for electricity infrastructure. Energy Res. Soc. Sci. 3, 152–160 (2014). doi: 10.1016/j.erss.2014.07.016 39. A. K. Chowdhury et al., Enabling a low-carbon electricity system for Southern Africa. Joule 6, 1826–1844 (2022). doi: 10.1016/j.joule.2022.06.030 40. International Renewable Energy Agency (IRENA), “Renewable power generation costs in 2020” (IRENA, 2021); https://www.irena.org/publications/2021/Jun/Renewable- Power-Costs-in-2020. 41. C. Callegari, A. Szklo, R. Schaeffer, Cost overruns and delays in energy megaprojects: How big is big enough? Energy Policy 114, 211–220 (2018). doi: 10.1016/j.enpol.2017.11.059 42. A. Ansar, B. Flyvbjerg, A. Budzier, D. Lunn, Should we build more large dams? The actual costs of hydropower megaproject development. Energy Policy 69, 43–56 (2014). doi: 10.1016/j.enpol.2013.10.069 43. G. M. Kondolf et al., Save the Mekong Delta from drowning. Science 376, 583–585 (2022). doi: 10.1126/science.abm5176; pmid: 35536906 44. J. J. Opperman et al., Balancing renewable energy and river resources by moving from individual assessments of Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E hydropower projects to energy system planning. Front. Environ. Sci. 10, 1036653 (2023). doi: 10.3389/ fenvs.2022.1036653 45. J. A. Harrison, Y. T. Prairie, S. Mercier‐Blais, C. Soued, Year‐2020 global distribution and pathways of reservoir methane and carbon dioxide emissions according to the greenhouse gas from reservoirs (G‐res) model. Glob. Biogeochem. Cycles 35, e2020GB006888 (2021). doi: 10.1029/2020GB006888 46. G. Wohlfahrt, E. Tomelleri, A. Hammerle, The albedo-climate penalty of hydropower reservoirs. Nat. Energy 6, 372–377 (2021). doi: 10.1038/s41560-021-00784-y; pmid: 33898056 47. C. Soued, J. A. Harrison, S. Mercier-Blais, Y. T. Prairie, Reservoir CO2 and CH4 emissions and their climate impact over the period 1900–2060. Nat. Geosci. 15, 700–705 (2022). doi: 10.1038/s41561-022-01004-2 48. R. M. Almeida et al., Reducing greenhouse gas emissions of Amazon hydropower with strategic dam planning. Nat. Commun. 10, 4281 (2019). doi: 10.1038/s41467-019-12179-5; pmid: 31537792 49. M. Li, N. He, Carbon intensity of global existing and future hydropower reservoirs. Renew. Sustain. Energy Rev. 162, 112433 (2022). doi: 10.1016/j.rser.2022.112433 50. C. Llamosas, B. K. Sovacool, The future of hydropower? A systematic review of the drivers, benefits and governance dynamics of transboundary dams. Renew. Sustain. Energy Rev. 137, 110495 (2021). doi: 10.1016/j.rser.2020.110495 51. P. M. Link, J. Scheffran, T. Ide, Conflict and cooperation in the water‐security nexus: A global comparative analysis of river basins under climate change. WIREs Water 3, 495–515 (2016). doi: 10.1002/wat2.1151 52. J. Yang et al., Quantifying the sustainability of water availability for the water-food-energy-ecosystem nexus in the Niger River basin. Earths Futur. 6, 1292–1310 (2018). doi: 10.1029/ 2018EF000923; pmid: 31032375 53. K. G. Wheeler et al., Exploring cooperative transboundary river management strategies for the Eastern Nile Basin. Water Resour. Res. 54, 9224–9254 (2018). doi: 10.1029/ 2017WR022149; pmid: 30774162 54. M. Zaniolo, M. Giuliani, S. Sinclair, P. Burlando, A. Castelletti, When timing matters-misdesigned dam filling impacts hydropower sustainability. Nat. Commun. 12, 3056 (2021). doi: 10.1038/s41467-021-23323-5; pmid: 34031413 55. A. Azzuni, C. Breyer, Definitions and dimensions of energy security: A literature review. WIREs Energy Environ. 7, e268 (2018). doi: 10.1002/wene.268 56. N. Saadi, A. Miketa, M. Howells, African Clean Energy Corridor: Regional integration to promote renewable energy fueled growth. Energy Res. Soc. Sci. 5, 130–132 (2015). doi: 10.1016/ j.erss.2014.12.020 57. K. J. Mach et al., Climate as a risk factor for armed conflict. Nature 571, 193–197 (2019). doi: 10.1038/s41586-019-1300-6; pmid: 31189956 58. A. Carlino, Data in support of “Declining cost of renewables and climate change curb the need for African hydropower expansion”. Zenodo (2023). doi: 10.5281/zenodo.7931050 ACKN OWLED GMEN TS Part of the research was developed while A.Car. was participating in the Young Scientists Summer Program at the International Institute for Applied Systems Analysis, Laxenburg. Funding: European Union’s Horizon 2020 grant 101003722 (GoNEXUS) (M.G., A.Cas.) Author contributions: Conceptualization: All authors contributed to the conceptualization; Methodology: A.Car., M.W., and M.G. A.Cas. designed the integrated modeling framework and the computational experiments. A.Car. performed data processing and ran the energy system model optimizations. C.J.C., S.S., W.T., and A.v.G. designed and performed computational experiments needed for this specific version of the African Hydropower Atlas; Visualization: A.Car.; Writing – original draft: A.Car., M.W., M.G., and A.Cas.; Writing – review and editing: All the authors reviewed and edited the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: The code and the instructions for replication of the computational experiments and to produce the figures supporting the results discussed in this paper are available at Zenodo (58). In particular, the repository contains the final energy demands, the updated version of the African Hydropower Atlas, the scripts for data processing and input preparation for the energy system model, the energy system model with instructions to run the optimizations, and the scripts to produce the figures. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf5848 Materials and Methods Supplementary Text Figs. S1 to S8 Table S1 References (59–82) Submitted 2 November 2022; accepted 1 July 2023 10.1126/science.adf5848 Carlino et al., Science 381, eadf5848 (2023) 11 August 2023 7 of 7
10.1126_science.adf3406
RES EARCH NEUROSCIENCE GluD1 binds GABA and controls inhibitory plasticity Laura Piot1, Christina Heroven2†‡, Simon Bossi1†, Joseph Zamith1, Tomas Malinauskas3, Chris Johnson2, Doris Wennagel1, David Stroebel1, Cécile Charrier1, A. Radu Aricescu2*, Laetitia Mony1*, Pierre Paoletti1* Fast synaptic neurotransmission in the vertebrate central nervous system relies primarily on ionotropic glutamate receptors (iGluRs), which drive neuronal excitation, and type A g-aminobutyric acid receptors (GABAARs), which are responsible for neuronal inhibition. However, the GluD1 receptor, an iGluR family member, is present at both excitatory and inhibitory synapses. Whether and how GluD1 activation may affect inhibitory neurotransmission is unknown. In this work, by using a combination of biochemical, structural, and functional analyses, we demonstrate that GluD1 binds GABA, a previously unknown feature of iGluRs. GluD1 activation produces long-lasting enhancement of GABAergic synaptic currents in the adult mouse hippocampus through a non-ionotropic mechanism that is dependent on trans-synaptic anchoring. The identification of GluD1 as a GABA receptor that controls inhibitory synaptic plasticity challenges the classical dichotomy between glutamatergic and GABAergic receptors. M ost excitatory neurotransmission in the central nervous system is mediated by the ionotropic glutamate receptor (iGluR) family members, which are classified as AMPA (GluA1 to 4), N- methyl-D-aspartate (NMDA) (GluN1, GluN2A to D, and GluN3A and B), kainate (GluK1 to 5), and delta (GluD1 and D2) receptors (1). At a typical excitatory synapse, glutamate released in the synaptic cleft binds postsynaptic iGluRs and drives opening of their cation-conductive ion channel, which results in membrane de- polarization. Although structurally related (2), GluD1 and GluD2 receptors depart from other iGluRs by their insensitivity to glutamate and inability to gate the ion channel pore upon ligand binding (3, 4). Instead, GluD receptors are thought to exert their physiological func- tion by binding D-serine (or glycine) (5–7) and through metabotropic (i.e., non-ionotropic) signaling (4, 8) [although ionotropic GluD sig- naling that is dependent on G protein–coupled receptors has also been suggested (9)]. This mechanism is best characterized for GluD2 receptors. Postsynaptic GluD2 binds cerebellin-1, a “molecular bridge” attached to presynaptic membrane proteins from the neurexin family, to form tripartite complexes that span the syn- aptic cleft and are involved in synapse forma- tion, maintenance, and plasticity (4, 10–13). Although GluD1 can also form triads with cerebellins and neurexins (8, 14), relatively little is known about GluD1 signaling and its functional roles. Expressed throughout the 1Institut de Biologie de l’ENS (IBENS), Ecole Normale Supérieure, Université PSL, CNRS, INSERM, F-75005 Paris, France. 2MRC Laboratory of Molecular Biology, Cambridge CB2 0QH, UK. 3Division of Structural Biology, Wellcome Centre for Human Genetics, University of Oxford, Oxford OX3 7BN, UK. *Corresponding author. Email: radu@mrc-lmb.cam.ac.uk (A.R.A.); laetitia.mony@ens.psl.eu (L.M.); pierre.paoletti@ens.psl.eu (P.P.) †These authors contributed equally to this work. ‡Present address: CAMS Oxford Institute, University of Oxford, Oxford OX3 7BN, UK. forebrain (15–17), GluD1 receptors participate in excitatory synaptogenesis (18, 19) and reg- ulate the balance between AMPA and NMDA receptors (20). GluD1 also specifies the forma- tion of inhibitory synapses in the neocortex (21–24). Results GluD1 receptors bind GABA Intrigued by the presence of GluD1 receptors at inhibitory postsynaptic sites (23, 24), which is unparalleled among iGluRs, we investigated the ligand-binding profile and activation of GluD1 receptors. Because agonist binding does not elicit currents in delta receptors, we first introduced in GluD1 the “Lurcher” A654T (Ala654→Thr) mutation, which renders GluD2 receptors constitutively active (25, 26). After the expression of GluD1-A654T mutant recep- tors into Xenopus oocytes, small but reliable constitutive inward currents were observed (fig. S1, A and B). Application of D-serine or glycine potentiated these currents (fig. S1, B and C), which contrasts with the current in- hibition observed on GluD2-Lurcher receptors (fig. S2, A to C) (5, 27). Because GluD1-A654T constitutive currents were small, we inserted an additional mutation, C645I (Cys645→Ile), to match the equivalent GluD2 M3 pore segment (fig. S1D). This double mutant GluD1-A654T- C645I, named GluD1-Lurcher hereafter, yielded larger constitutive currents (fig. S1E) that were also potentiated by D-serine and glycine (Fig. 1, A to C) while displaying almost no sensitivity to glutamate (Fig. 1C and fig. S3, A to C). Agonist dose-response curves on GluD1-Lurcher revealed D-serine potency in the high micromolar range [median effective concentration (EC50) of 320 ± 10 mM (n = 13 to 17 cells); Fig. 1C and table S1], similar to that measured on GluD2-Lurcher receptors [median inhibitory concentration (IC50) of 260 ± 10 mM (n = 5 or 6 cells); fig. S2, A to C, and (5, 27, 28)]. Glycine sensitivity appeared rightward shifted, in the low milli- molar range [EC50 of 3.4 ± 0.2 mM (n = 9 cells); Fig. 1C and table S1], reminiscent of the glycine sensitivity of GluD2-Lurcher recep- tors [IC50 = 0.75 ± 0.05 mM (n = 10 cells); fig. S2C]. We then tested the ability of GluD1 re- ceptors to sense g-aminobutyric acid (GABA). GABA robustly potentiated GluD1-Lurcher cur- rents with an apparent affinity in the low milli- molar range [EC50 = 3.0 ± 0.2 mM (n = 15 to 17 cells)], lower than that of D-serine but sim- ilar to that of glycine, and with equal efficacy (Fig. 1, B and C; see also fig. S1, B and C, and table S1). This GABA effect was displaced by D-serine (fig. S3, D to F). By contrast, GABA was essentially ineffective on GluD2-Lurcher recep- tors, producing modest current inhibition and limited D-serine displacement only at high con- centrations (fig. S2, D to F). Mutating conserved iGluR agonist-binding residues (1) in the GluD1 ligand-binding domain (LBD) fully abolished responses to GABA (figs. S2, G to I, and S3, G to I), revealing the primary role of the GluD1 LBD in GABA sensing. Application of GABA (10 mM) onto NMDA (fig. S4, A to D) and AMPA (fig. S4, E to H) receptors triggered negligible effects, which is consistent with insensitivity to GABA. We next aimed to confirm these observations in GluD1 receptors devoid of the Lurcher muta- tions. We used voltage clamp fluorometry (VCF), a technique that enables detection of ligand (or voltage)–induced conformational changes in membrane proteins (29). We designed a series of GluD1 receptors with cysteines engineered at various positions in the LBDs for attachment of the environment-sensitive fluorescent probe Alexa488 and screened for potential VCF fluo- rescence changes (DFs) upon ligand addition (Fig. 1D). Several positions yielded robust DFs upon D-serine (or glycine) application. Posi- tion E525 (E, Glu) in the structurally sensitive D1-D1 dimer interface (27) produced particu- larly strong signals (table S2). We verified that the E525C (Glu525→Cys) mutation and the flu- orescent probe attachment had minimal ef- fects on the properties of the receptor (fig. S5). Application of GABA on labeled GluD1- E525C receptors also triggered marked DFs, which revealed the ability of GABA to bind and elicit conformational changes at (non-Lurcher) GluD1 receptors (Fig. 1, D to F, and fig. S6, A to C). Glutamate application had hardly any effect (Fig. 1, E and F). No fluorescent signals were detected upon D-serine, glycine, or GABA ad- dition onto binding-deficient GluD1 receptors (Fig. 1, E and F, and fig. S6, D to F), which demonstrates that ligand binding at the GluD1 LBD is the critical initial molecular event. VCF-based dose-response curves showed agonist apparent affinities in the sub- to low-millimolar range, with GABA sensitivity similar to that of glycine but lower than that of D-serine (fig. S6, B and C), which thus corroborates results ob- tained on GluD1 Lurcher receptors. Thermal Piot et al., Science 382, 1389–1394 (2023) 22 December 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. GABA binds and triggers conformational changes at GluD1 recep- tors. (A) Schematic of the two-electrode voltage clamp technique performed on Xenopus oocytes that express GluD1-Lurcher receptors. TMD, transmembrane domain. (B) D-serine and GABA potentiate GluD1-Lurcher constitutive currents. The pore blocker pentamidine (Penta, 100 mM) was used to estimate the amount of constitutive current. Dashed lines represent the constitutive Lurcher currents. (C) Dose-response curves for GABA, D-serine, glycine, and glutamate on GluD1- Lurcher receptors. Values and sample numbers are provided in table S1. Error bars indicate SD. (D) Schematic of VCF performed on Xenopus oocytes that express GluD1*-E525C labeled with Alexa488-maleimide. GluD1* is the receptor containing the C721S mutation (Cys721→Ser). (E) VCF signals of GluD1*-E525C and GluD1*-E525C-R526K-D742A receptors upon application of D-serine, glycine, GABA, and glutamate (10 mM each). R526K, Arg526→Lys; D742A, Asp742→Ala. (F) DF amplitudes triggered by 10-mM ligand application on oocytes that express either GluD1*-E525C or GluD1*-E525C-R526K-D742A receptors and on non- injected oocytes. The center line represents the median, box limits are upper and lower quartiles, and whiskers are minimum and maximum values. Values and sample numbers are listed in table S5. ***P < 0.001, and n.s. is not significant (Dunn’s post hoc test). Detailed replicate numbers and the full statistical analysis are provided in table S6. shift assay and isothermal titration calorimetry (ITC) experiments on isolated LBDs confirmed direct binding of GABA (and D-serine, but not glutamate) to the GluD1 LBD with a binding affinity of 2 mM (Fig. 2, A and B; fig. S7, A and B; and table S3), which is consistent with our results on full-length GluD1 receptors. Con- versely, GluD2 LBDs bound D-serine but not GABA (fig. S7, C to F). Therefore, GABA ap- pears to be an active ligand at GluD1, but not GluD2, receptors. Structure of the GABA binding site To understand how GluD1 interacts with GABA, we determined the crystal structure of the GluD1 LBD–GABA complex using x-ray crystallography (Fig. 2, C and D; fig. S8, A and B; and table S4). The GluD1 LBD–GABA complex formed well- ordered crystals that diffracted to 1.9-Å reso- lution, which allowed for the unambiguous identification of GABA electron density (Fig. 2D). GABA binds the central LBD interlobe cleft and, typical of iGluR agonists (1, 30), forms multiple interactions with both the upper and lower lobes (Fig. 2, D and E). In particular, the guanidinium group of GluD1-R526 (R, Arg) engages in a bidentate salt bridge with the carboxyl group of GABA, whereas the g-amino group forms a dual electrostatic interaction with the carboxylates of GluD1-E446 (upper lobe) and GluD1-D742 (lower lobe) (D, Asp). Disrupting these interactions prevented GABA action on GluD1 receptors (fig. S3, G to I). Thus, charged interactions that include the strictly conserved bond between the arginine side chain of LBD helix D and the ligand a-carboxylate (1, 30) are major anchors for GABA binding on GluD1. In addition, three aromatic tyrosine (Y) residues line the GABA binding pocket, including GluD1-Y492, which acts as a lid on top of the GABA molecule (Fig. 2, D and E). For comparison, we also solved the structure of the GluD1 LBD in the apo state and in the presence of D-serine (Fig. 2C; fig. S8, C to H; and table S4). The apo GluD1 LBD crystal form contained three dimers in the asymmetric unit and displayed open cleft conformations (Fig. 2C and fig. S9, A to C), which is reminis- cent of other apo iGluR LBD structures (1, 30). By contrast, the GABA-bound GluD1 LBD crystallized as an asymmetric dimer, in which LBDs adopt closed-cleft conformations, albeit the closure angles differed between the two protomers (17.3° and 10.5°, respectively, rel- ative to chain E of the apo state; Fig. 2C and fig. S8, A to C). The GluD1 LBD–D-serine com- plex crystal diffracted to 1.63-Å resolution and contained one homodimer per asymmetric unit (table S4). Only one monomer had D-serine bound, and it adopted a closed-cleft confor- mation identical to that observed with GABA (closure of 17.3° relative to chain E of the apo state; fig. S8, D to F). The D-serine interactions with GluD1 LBD lobes again involve both ionic and hydrogen bonding (Fig. 2F and fig. S8F). Overall, the ligand-binding pose is notably sim- ilar to that described for D-serine bound to the GluD2 receptor LBD (5, 31) (fig. S8). Molecular determinants of GABA selectivity To examine the molecular basis of GluD1 GABA binding and selectivity, we performed structure- based mutagenesis experiments. Substituting the LBD upper lobe residue E446 to Ser (S), Gln (Q), or Asp (N) abolished GABA effects on GluD1-Lurcher currents, but D-serine remained active (Fig. 3, A to C, and fig. S10). The more conservative E446D (Glu446→Asp) mutant was not as discriminating and retained sensitivity to both ligands, although with reduced affinity (Fig. 3C, fig. S10, and table S1). These results are consistent with E446 making ionic inter- actions with GABA but not D-serine, which thus identifies E446 as pivotal for GABA recognition Piot et al., Science 382, 1389–1394 (2023) 22 December 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E A GluD1LBD vs D-serine B GluD1LBD vs GABA C r e w o p l a i t n e r e f f i d s s e c x E t a e h s s e c x E 6.2 6.0 5.8 5.6 5.4 ) s / l a c µ ( ) t n a t c e n j i l o m / l a c k ( 0.0 -0.2 -0.4 -0.6 D R526 6.2 6.1 6.0 5.9 5.8 5.7 0 10 20 30 40 50 60 Time (min) 0 10 20 30 40 50 60 Time (min) KD = 248 µM 0.10 0.00 -0.01 -0.02 -0.03 -0.04 -0.05 KD = 2 mM 0 10 20 30 40 50 Molar ratio N 0 100 50 Molar ratio N 150 Upper lobe Y492 E446 17.3° 17.3° GluD1LBD + GABA GluD1LBD + D-Serine GABA 17.3° GluD1LBD apo D-Serine 17.3° E Asp 742 CA F Trp 741 Trp 741 Ala 686 Val 687 Asp 742 CA 2.9 3.0 2.9 2.7 3.0 2.9 GABA 2.8 T521 Lower lobe D742 Glu 446 CA CA Ala 519 Tyr 492 GABA CA Ala 686 Tyr 539 D-Ser Thr 521 CA Ala 519 CA Ile 520 Ile 520 Arg 526 CA Thr 521 Tyr 492 CA CA Arg 526 Fig. 2. Structures of the GABA-GluD1 and D-serine–GluD1 LBD complexes. (A and B) ITC analysis of D-serine (A) and GABA (B) binding to the isolated GluD1 LBD; values are listed in table S3. KD, dissociation constant. (C) X-ray crystal structures of the GluD1 LBD in complex with GABA (left, blue) and D-serine (right, cyan) superimposed with the apo GluD1 LBD structure (tan). The extent of GluD1 LBD clamshell closure upon ligand binding is indicated. (D) Close-up view of the GABA binding site, which shows the 2mFo-DFc electron density map contoured at 1.0s. Distances (in Å) between interacting atoms are indicated. (E and F) LigPlot diagrams of the GluD1-GABA (E) and GluD1–D-serine (F) complexes. CA, a-carbon atom. Fig. 3. Molecular deter- minants of GABA recog- nition by GluD1 receptors. (A) Schematic of a GluD1-Lurcher recep- tor that harbors the E446Q (Glu446→Gln) mutation. (B) The E446Q mutation eliminates the sensitivity of GluD1- Lurcher to GABA but not to D-serine. Dashed lines represent the constitutive Lurcher currents. (C) GABA dose-response curves of various GluD1- Lurcher-E446 mutants. Error bars indicate SD. (D) A schematic of the GluN1-Q405E-Q536Y/ GluN2A NMDA receptor mutant is shown at the top. An alignment of the GluD1 and GluN1 LBD sequences is shown at the bottom. Q405E, Gln405→Glu; Q536Y, Gln536→Tyr. (E) Mutations Q405E and Q536Y in the GluN1 LBD confer GABA sensitivity to GluN1/GluN2A NMDA receptors. Current traces elicited by successive applications of glutamate plus glycine, GABA alone, and glutamate plus GABA are shown. (F) GABA and glycine (coagonist) dose- response curves (in 1 mM glutamate) of GluN1-Q405E-Q536Y/GluN2A and wild-type (WT) GluN1/GluN2A receptors. The latter receptors are insensitive to GABA (10 mM). Error bars indicate SD. Values and sample numbers are provided in table S1. Piot et al., Science 382, 1389–1394 (2023) 22 December 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E A IUE E15.5 - + Acute slice Adult mice Control ShGluD1 CA1 slm CA2 CA3 D Recording B Stimulation N H P G 1 D u G l e g r e M Control GluD1# E 5 9 - D S P 1 D u G l e g r e M i n o i t a c o s s a f o e g a t n e c r e P 80 60 40 20 0 *** GPHN PSD-95 Post HFS Baseline 100pA 25ms * ) % ( e d u t i l p m a C S P I e v i t a e R l 150 125 100 75 50 100 pA 25 ms HFS 50 pA 25 ms n.s Control ShGluD1 + GluD1# 0 5 15 10 Time (min) 20 25 Control GluD1#-V617R 50 pA HFS 25 ms 100 pA 25 ms n.s. Control ShGluD1 + GluD1#-V617R 0 5 10 15 Time (min) 20 25 ) % ( e d u t i l p m a C S P I e v i t a e R l 150 125 100 75 50 H ) % ( e d u t i l p m a C S P I e v i t a e R l 150 125 100 75 50 Control GluD1#-R526K 50 pA 25 ms HFS 50 pA 25 ms ** Control ShGluD1 + GluD1#-R526K 0 5 10 15 Time (min) 20 25 Control GluD1#-E446S GABA D-serine or glycine 100 pA 25 ms HFS 100 pA 25 ms **** Control ShGluD1 + GluD1#-E446S 0 5 10 15 Time (min) 20 25 Control ShGluD1 5 10 15 Time (min) 20 F Control GluD1#-R341A-W343A G 50 pA 25 ms HFS 50 pA 25 ms * Control ShGluD1 + GluD1#-R341A-W343A 0 5 10 15 Time (min) 20 25 ) % ( e d u t i l p m a C S P I e v i t a e R l 150 125 100 75 50 * n.s. ** * n.s. **** J l o r t n o C l 1 D u G h S l o r t n o C l # 1 D u G + 1 D u G h S l l o r t n o C l 1 D u G h S l o r t n o C K 6 2 5 R - # 1 D u G + l l # 1 D u G + 1 D u G h S l l o r t n o C l o r t n o C R 7 1 6 V - # 1 D u G + l l 1 D u G h S A 3 4 3 W A 1 4 3 R - - S 6 4 4 E - 1 D u G + l # l 1 D u G h S GABA D-serine or glycine SR KO 50 pA 25 ms HFS Serine Racemase KO ) % ( e d u t i l p m a C S P I e v i t a e R l 150 100 50 0 0 10 20 Time (min) 30 SR KO 150 125 100 75 50 ) % ( e d u t i l p m a C S P I e v i t a e R l 100pA 25ms 150 125 HFS C ) % ( e d u t i l p m a C S P I e v i t a e R l 100 75 50 0 ) % ( e d u t i l p m a C S P I e v i t a e R l 150 125 100 75 50 I ) % ( e d u t i l p m a C S P I e v i t a e R l 200 150 100 50 0 Fig. 4. GluD1 receptors mediate GABAergic synaptic plasticity. (A) A schematic of in utero electroporation (IUE) for cell-specific labeling and manipulation of GluD1 expression in the mouse hippocampus performed at embryonic day 15.5 (E15.5) is shown on the left. A confocal image of the hippocampus of a P21 mouse electroporated with tdTomato that shows selective targeting of CA1 pyramidal neurons is shown on the right. Imaging and whole-cell patch-clamp recordings were performed at CA1 synapses in the SLM, a region enriched in GluD1 receptors (15–17). Scale bar, 0.5 mm. (B) GluD1 preferentially associates with inhibitory SLM synapses. Colocalization images are shown on the left. The white arrowheads highlight the association between GluD1 and the indicated synaptic markers. A quantification is shown on the right. The center line represents the median, box limits are upper and lower quartiles, and whiskers are minimum and maximum values. For gephyrin (GPHN), n = 20 dendrites, and for postsynaptic density protein 95 (PSD-95), n = 18 dendrites; ***P < 0.001 (two-sided unpaired Student’s t test). Scale bars, 2 mm. (C) HFS potentiates IPSCs in control neurons but not in neurons that express short hairpin RNA downregulating GluD1 expression (ShGluD1). For each condition, n = 8 cells. Representative IPSC traces before (black) and after (green) HFS are shown at the top. (D) Molecular replacement experiments. Reexpression of WT GluD1 rescues HFS-induced plasticity. n = 7 cells for control, and n = 8 cells for ShGluD1 plus GluD1# (the # indicates the shRNA-resistant GluD1 construct). (E to H) Ligand binding–deficient (E) (n = 9 cells versus n = 10 cells for control) and cerebellin binding–deficient (F) (n = 8 cells versus n = 10 cells for control) GluD1 receptors do not rescue HFS-induced inhibitory plasticity in neurons in which GluD1 is knocked down, but ion channel flux–deficient (G) (n = 9 cells versus n = 8 cells for control) GluD1 receptors do. GluD1-E446S (H), which allows binding of glycine and D-serine but not of GABA, does not rescue HFS plasticity (n = 9 cells versus n = 8 cells for control). R341A, Arg341→Ala; V617R, Val617→Arg; W343A, Trp343→Ala. (I) Pooled data of molecular replacement experimental Piot et al., Science 382, 1389–1394 (2023) 22 December 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E results on HFS-induced IPSC amplitude. The control group is nonelectroporated neurons from the same electroporated mouse. (J) HFS-induced inhibitory plasticity is present in serine racemase knockout (SR KO) mice (n = 8 cells) (left). A quantification of HFS-induced plasticity in SR KO mice is shown on the right. Error bars indicate SEM. Detailed replicate numbers and the full statistical analysis are provided in in table S6. For (C) to (I), *P < 0.05, **P < 0.01, ***P < 0.0001, and n.s. is not significant (two-sided Mann-Whitney test); error bars indicate SEM. by GluD1. The importance of this residue as a key GABA anchor was reinforced by gain-of-function experiments performed on (GABA-insensitive) NMDA receptors. In the glycine-binding GluN1 subunit, the equivalent residue of GluD1-E446 is a glutamine (Q405) (Fig. 3D and fig. S11). Introduction of the single-point mutation Q405E (Gln405→Glu) into the GluN1 LBD allowed direct activation of GluN1/GluN2A NMDA receptors by coapplication of glutamate and GABA (in replacement of glycine; fig. S12, A to C). The gating efficacy of GABA at NMDA receptors was further enhanced by introducing a second mutation in GluN1 [mutant GluN1-Q405E- Q536Y (Gln536→Tyr)] to better mimic the GluD1 GABA binding environment (Fig. 3, D and E, and fig. S12C). On GluN1-Q405E-Q536Y/GluN2A receptors, the efficacy of GABA as a coagonist reached >50% of that of glycine (Fig. 3F), which points to E446 and Y539 as critical determi- nants of GABA recognition in GluD1 receptors. E446 and Y539 are conserved in GluD2 recep- tors (fig. S11), which suggests that other GluD1 determinants contribute to GABA selectivity among delta receptors. GABA binding to GluD1 controls inhibitory synaptic strength We then assessed the impact of GluD1 activation on synaptic transmission at mature GABAergic synapses. We focused on the stratum lacunosum- moleculare (SLM) of the mouse hippocampus, a region highly enriched in GluD1 receptors (15–17) (Fig. 4A). Confocal microscopy imaging revealed that GluD1 preferentially accumu- lates at inhibitory synapses rather than at excitatory synapses (Fig. 4B), as previously ob- served in distal dendrites of cortical pyramidal neurons (23, 24). We recorded GABAergic in- hibitory postsynaptic currents (IPSCs) from CA1 pyramidal cells while performing electrical stim- ulations in the SLM (fig. S13A). Application of the GluD1 agonist D-serine potentiated the IPSC amplitudes [EC50 of 180 ± 0.05 mM (n = 5 to 16 cells); fig. S13, B to D]. The continuous inclu- sion of D-aminophosphovalerate (D-APV) in these experiments ruled out a direct activation of NMDA receptors by D-serine as an underlying mech- anism. Application of 5,7-dichlorokynurenic acid (DCKA), a ligand that, like D-serine, binds iGluR glycine sites (7) and potentiates GluD1- Lurcher receptors (fig. S14), also produced robust enhancement of IPSC amplitude (fig. S13E). Short hairpin RNA (shRNA)–based down-regulation of GluD1 expression following hippocampus-targeted in utero electropora- tion eliminated the D-serine– and DCKA-induced potentiation (fig. S13, B and E). It furthermore decreased the density of SLM inhibitory syn- apses (fig. S13F), as previously shown in the cortex (23). This indicates both developmen- tal and acute effects of GluD1 signaling on GABAergic transmission at SLM CA1 syn- apses. We next induced short bursts of high- frequency stimulation (HFS) to promote the release of endogenous GABA from presynaptic inhibitory terminals (32). After HFS, IPSC am- plitude increased by ~30% in control neurons but not in neurons in which GluD1 was knocked down, an effect that persisted throughout the recording (15 min; Fig. 4C) and likely involved a postsynaptic mechanism (fig. S13G). A mo- lecular replacement strategy with mutant GluD1 receptors deficient in distinct signaling modali- ties (23) revealed that ligand binding and inter- action with the extracellular scaffolding protein cerebellin, but not ion flux through GluD1 chan- nels, are required for GluD1 to mediate enhance- ment of inhibitory neurotransmission (Fig. 4, D to I, and fig. S13H). Preventing binding of GABA, but not of D-serine or glycine, with the GluD1-E446S (Glu446→Ser) mutant abolished HFS-induced plasticity (Fig. 4, H and I), whereas inhibitory plasticity was conserved in mice de- ficient in D-serine signaling (33) (Fig. 4J). Thus, activation of GluD1 receptors at SLM CA1 hippo- campal GABAergic synapses induces prolonged potentiation of inhibitory synaptic strength through a non-ionotropic GABA-induced mech- anism that relies on trans-synaptic anchoring. Discussion Our study unveils that, in the SLM of the mouse hippocampus, GluD1 receptors operate at GABAergic synapses as receptors that bind GABA and control inhibitory synaptic plasticity. We pro- pose that synaptically released GABA, together with cerebellin, binds to and triggers conforma- tional changes of postsynaptic GluD1 receptors, which activates intracellular signaling path- ways, leading to an increase in the number and/ or activity of type A GABA receptors (GABAARs). This GluD1-driven mechanism of inhibitory plasticity, which is embedded in the inhibitory synapse itself, departs from previously des- cribed mechanisms of inhibitory plasticity that commonly rely on cross-talk with neighboring excitatory (glutamatergic) synapses (34). By contrast, the non-ionotropic, trans-synaptic, and ligand-induced mechanism echoes the GluD2- induced plasticity at glutamatergic parallel fiber–Purkinje cell synapses in the cerebellum, where GluD2 activation by extracellular D-serine regulates AMPA receptor content (4, 6, 35, 36). The high (10 mM) concentrations of neuro- transmitter reached at fast chemical synapses after vesicular release (37–39) indicate substan- tial activation of GluD1 receptors by GABA despite their relatively low affinity for this agonist. The detailed mechanisms that couple GluD1 activation to the modulation of GABAAR activity, however, remain to be deciphered. The question of whether this previously unknown form of GABAergic synaptic plasticity interacts with other forms of inhibitory plasticity (34) also remains open. Because GluD1 receptors are broadly expressed in the forebrain (15–17), including at inhibitory synapses in the neo- cortex (23), GABA signaling through GluD1 receptors likely represents a general mechanism that extends the computational rules of inhib- itory plasticity with consequences on neuronal circuit function. In humans, GluD1 mutations are associated with susceptibility to autism and schizophrenia (4, 40) as well as major depressive disorders (41). With their dual ability to reside at excitatory and inhibitory synapses, GluD1 receptors are not only specially equipped to act as powerful reg- ulators of synaptic circuits but are also vulnerable nodes of excitation-inhibition imbalance during neuropsychiatric disorders. In that context, our finding that GluD1 receptors are molecular machines with hybrid features—functionally GABAergic but structurally glutamatergic— opens fresh perspectives on GluD1-targeted neuropharmacology. REFERENCES AND NOTES 1. K. B. Hansen et al., Pharmacol. Rev. 73, 298–487 (2021). 2. A. P. Burada, R. Vinnakota, J. Kumar, Nat. Struct. Mol. Biol. 27, 84–91 (2020). 3. S. M. Schmid, S. Kott, C. Sager, T. Huelsken, M. Hollmann, Proc. Natl. Acad. Sci. U.S.A. 106, 10320–10325 (2009). 4. M. Yuzaki, A. R. Aricescu, Trends Neurosci. 40, 138–150 (2017). 5. P. Naur et al., Proc. Natl. Acad. Sci. U.S.A. 104, 14116–14121 (2007). 6. W. Kakegawa et al., Nat. Neurosci. 14, 603–611 (2011). 7. D. Stroebel, L. Mony, P. Paoletti, Neuropharmacology 193, 108631 (2021). 8. M. Fossati, C. Charrier, Curr. Opin. Neurobiol. 66, 85–92 (2021). 9. A. P. Burada, R. Vinnakota, B. Lambolez, L. Tricoire, J. Kumar, Neuropharmacology 196, 108683 (2021). 10. T. Uemura et al., Cell 141, 1068–1079 (2010). 11. K. Matsuda et al., Science 328, 363–368 (2010). 12. J. Elegheert et al., Science 353, 295–299 (2016). 13. J. Dai, K. Liakath-Ali, S. R. Golf, T. C. Südhof, eLife 11, e78649 (2022). 14. P. C. Andrews, S. M. Dravid, Neuropharmacology 192, 108587 (2021). 15. K. Konno et al., J. Neurosci. 34, 7412–7424 (2014). 16. R. Hepp et al., Brain Struct. Funct. 220, 2797–2815 (2015). 17. C. Nakamoto et al., J. Comp. Neurol. 528, 1003–1027 (2020). 18. S. C. Gupta et al., Neuropharmacology 93, 274–284 (2015). 19. W. Tao, J. Díaz-Alonso, N. Sheng, R. A. Nicoll, Proc. Natl. Acad. Sci. U.S.A. 115, E5373–E5381 (2018). 20. J. Dai, C. Patzke, K. Liakath-Ali, E. Seigneur, T. C. Südhof, Nature 595, 261–265 (2021). 21. K. Ryu, M. Yokoyama, M. Yamashita, T. Hirano, Biochem. Biophys. Res. Commun. 417, 157–161 (2012). 22. M. Yasumura et al., J. Neurochem. 121, 705–716 (2012). 23. M. Fossati et al., Neuron 104, 1081–1094.e7 (2019). 24. E. Favuzzi et al., Science 363, 413–417 (2019). 25. J. Zuo et al., Nature 388, 769–773 (1997). 26. L. P. Wollmuth et al., J. Neurosci. 20, 5973–5980 (2000). 27. K. B. Hansen et al., J. Neurosci. 29, 907–917 (2009). Piot et al., Science 382, 1389–1394 (2023) 22 December 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 28. D. Tapken et al., Sci. Rep. 7, 46145 (2017). 29. L. M. Mannuzzu, M. M. Moronne, E. Y. Isacoff, Science 271, 213–216 (1996). 30. J. Pøhlsgaard, K. Frydenvang, U. Madsen, J. S. Kastrup, Neuropharmacology 60, 135–150 (2011). 31. A. S. Kristensen et al., Mol. Pharmacol. 89, 253–262 (2016). 32. Y. Komatsu, J. Neurosci. 14, 6488–6499 (1994). 33. A. C. Basu et al., Mol. Psychiatry 14, 719–727 (2009). 34. C. Q. Chiu, A. Barberis, M. J. Higley, Nat. Rev. Neurosci. 20, 272–281 (2019). 35. W. Kakegawa, K. Kohda, M. Yuzaki, J. Physiol. 584, 89–96 (2007). 36. K. Kohda et al., Proc. Natl. Acad. Sci. U.S.A. 110, E948–E957 (2013). 37. B. Barbour, M. Häusser, Trends Neurosci. 20, 377–384 (1997). 38. J. D. Clements, R. A. Lester, G. Tong, C. E. Jahr, G. L. Westbrook, Science 258, 1498–1501 (1992). 39. A. Reiner, J. Levitz, Neuron 98, 1080–1098 (2018). 40. A. J. Griswold et al., Hum. Mol. Genet. 21, 3513–3523 (2012). 41. P. Muglia et al., Mol. Psychiatry 15, 589–601 (2010). ACKN OW LEDG MEN TS We thank L. Tricoire [Institute of Biology Paris Seine (IBPS), Sorbonne Université, Paris, France] for the gift of the GluD1 and GluD2 plasmids, D. B. Arnold (University of Southern California, USA) for plasmids available through Addgene, D. Bellini and F. Gorrec [MRC Laboratory of Molecular Biology (MRC-LMB), UK] for access to the crystallization facility and x-ray data collection, and J. Grimmett and T. Darling (MRC-LMB, UK) for support with scientific computing. We thank P.-J. Corringer (Institut Pasteur, Paris, France), S. Pless (University of Copenhagen, Denmark), and members of their teams for helpful discussions and advice on VCF, as well as the Institut de Biologie de l’École Normale Supérieure (IBENS) FabLab for help with 3D printing of the VCF perfusion chamber. We also thank N. Rebola [Institut du Cerveau (ICM), Paris] for the serine racemase knockout mice. Funding: This work was supported by the French government (Investissements d’Avenir ANR-10-LABX-54 MEMOLIFE and ANR-11-IDEX-0001-02 PSL* Research University), Sorbonne University (fellowships to L.P.), the Fondation pour la Recherche Médicale (FRM) (fellowship no. FDT202012010605 to L.P.), the European Research Council (ERC Advanced Grant no. 693021 to P.P. and ERC Starting Grant no. 803704 to C.C.), and the UK Medical Research Council (grants MR/L009609/1 and MC UP 1201/15 to A.R.A.). Author contributions: Experiments on Xenopus oocytes (two-electrode voltage clamp and VCF): L.P.; VCF set-up building and operation: L.M. and L.P.; Crystallography experiments analysis and design: C.H., T.M., and A.R.A.; Biochemical experiments (ITC) analysis and design: C.J. and C.H.; In utero electroporation experiments and design, confocal imaging, and analysis: J.Z. and C.C.; Western blot experiments: D.W.; Brain slice electrophysiological experiments: S.B.; Study design and data analysis: L.M., D.S., P.P.; Writing – original draft, review, and editing: L.P., C.C., A.R.A., L.M., P.P. Competing interests: A.R.A. is an employee of BioNTech UK Ltd. and has equity interests in BioNTech SE. All other authors declare that they have no competing interests. Data and materials availability: All data are available in the main text or the supplementary materials. Structure factors and coordinates of apo, D-serine–bound, and GABA-bound GluD1 LBDs were deposited in the Protein Data Bank (PDB) under IDs 8BLJ, 8BN2, and 8BN5, respectively. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf3406 Materials and Methods Figs. S1 to S14 Tables S1 to S6 References (42–69) MDAR Reproducibility Checklist Submitted 20 November 2022; accepted 25 October 2023 Published online 7 December 2023 10.1126/science.adf3406 Piot et al., Science 382, 1389–1394 (2023) 22 December 2023 6 of 6
10.1126_science.adf8042
transfer interactions during C–H activation by metal complexes. Using time-resolved x-ray absorption spectroscopy (XAS) at the metal L-edge (11, 25–30), we probe the short-lived reaction intermediates from the vantage point of the reactive metal site to interrogate the decisive charge-transfer interactions that deter- mine the overall reaction. In two ultraviolet (UV)–pump and x-ray–probe experiments at the Swiss Free Electron Laser facility (SwissFEL) and the Swiss Light Source synchrotron radi- ation facility (SLS), we track s-complex forma- tion and oxidative addition using CpRh(CO)2 (where Cp is cyclopentadienyl) (10, 18–20) in octane solution. The time-resolved Rh L-edge absorption spectra were recorded by collecting the x-ray fluorescence as a function of incident x-ray photon energy around the Rh L3 absorption edge (Fig. 1C; see supplementary materials for experimental details). As is the case for other 4d transition metal complexes (27, 31, 32), the Rh L3-edge transitions can be assigned to ex- citations of Rh 2p core electrons to unoccupied molecular orbitals (see Fig. 1D). Changes in transition energies reflect changes in orbital energies, whereas oscillator strengths vary with the degree to which Rh 4d and ligand orbitals hybridize. In combination with our calculations, the data provide direct access to back-and-forth charge-transfer interactions along the C–H activation reaction trajectory at the level of individual orbitals. Time-resolved XAS of C–H activation The steady-state Rh L3-edge absorption spec- trum of CpRh(CO)2 shown in Fig. 1D exhibits a peak at a photon energy of ~3006 eV that results from excitation of Rh 2p core elec- trons into the lowest unoccupied molecular orbital (LUMO), the empty 4d-derived orbital of the Rh(I) d8 ground state configuration. The second peak at ~3007.5 eV is assigned to transitions of Rh 2p electrons into unoccupied orbitals of mainly CO and/or Cp ligand char- acter. Through metal-ligand back-donation, these ligand-derived orbitals acquire Rh 4d character and become accessible by the Rh 2p→d dipole transitions in L3-edge XAS (33). Upon laser excitation, as seen in the dif- ference spectrum recorded at a pump-probe time delay of 250 fs, a pre-edge peak appears at ~3002.5 eV together with substantial bleaching RES EARCH ULTRAFAST DYNAMICS Tracking C–H activation with orbital resolution Raphael M. Jay1*†, Ambar Banerjee1*†, Torsten Leitner1‡, Ru-Pan Wang2, Jessica Harich2, Robert Stefanuik1, Hampus Wikmark1§, Michael R. Coates3, Emma V. Beale4, Victoria Kabanova4¶, Abdullah Kahraman4#**, Anna Wach4,5, Dmitry Ozerov4, Christopher Arrell4, Philip J. M. Johnson4, Camelia N. Borca4, Claudio Cirelli4, Camila Bacellar4, Christopher Milne6, Nils Huse2, Grigory Smolentsev4, Thomas Huthwelker4, Michael Odelius3, Philippe Wernet1* Transition metal reactivity toward carbon–hydrogen (C–H) bonds hinges on the interplay of electron donation and withdrawal at the metal center. Manipulating this reactivity in a controlled way is difficult because the hypothesized metal-alkane charge-transfer interactions are challenging to access experimentally. Using time-resolved x-ray spectroscopy, we track the charge-transfer interactions during C–H activation of octane by a cyclopentadienyl rhodium carbonyl complex. Changes in oxidation state as well as valence-orbital energies and character emerge in the data on a femtosecond to nanosecond timescale. The x-ray spectroscopic signatures reflect how alkane-to-metal donation determines metal-alkane complex stability and how metal-to-alkane back-donation facilitates C–H bond cleavage by oxidative addition. The ability to dissect charge-transfer interactions on an orbital level provides opportunities for manipulating C–H reactivity at transition metals. density from the occupied C–H s-orbital into unoccupied metal d-orbitals concomitant with back-donation from occupied metal d-orbitals into the unoccupied antibonding C–H s*-orbital (21–24) (similar to, albeit substantially weaker than, metal-carbonyl bonds, as illustrated in Fig. 1B). Both types of interactions simulta- neously enhance metal-alkane bonding and weaken the alkane C–H bond. Because it is the balance of back-and-forth charge-transfer via different orbitals that determines whether a s-complex ultimately proceeds to C–H bond cleavage, dissecting individual charge-transfer interactions could provide orbital-based design principles as a guide for catalyst development. Experimentally, time-resolved infrared (IR) spectroscopy has been instrumental in iden- tifying reaction intermediates in C–H activa- tion (17) by probing shifts in infrared marker modes of spectator ligands. Such shifts are the result of changes in spectator-ligand bond strengths induced by changes in the integrated charge-transfer interactions in the complex. Sepa- rately accessing donation and back-donation to and from the metal in a s-complex, however, would be a way to experimentally correlate individual orbital interactions with reactivity toward C–H bond cleavage (7). In this work, we demonstrate a distinct way to experimentally evaluate metal-ligand charge- T he transformation of saturated hydro- carbons under mild conditions into more valuable products constitutes a long- standing challenge in chemistry (1–4). Photoinitiated reactions of transition metal carbonyl complexes with alkanes have long served as fruitful model systems (5, 6), providing detailed insights into the cleavage mechanism of strong C–H bonds at a metal center (1, 2, 4, 7). In these systems, photo- induced ligand loss is known to create a highly reactive species with an undercoordinated and electron-deficient metal center (Fig. 1A). The metal then rapidly binds an alkane from solution to form a s-complex, in which the metal coordinates to one or more C–H s-bonds. Ultimately, metal insertion between C and H atoms breaks the C–H bond to form a metal alkyl hydride product. The s-complex inter- mediates have been extensively studied over the past several decades to probe their molecu- lar structure and mechanistic role (8–20). Quantum chemical calculations, in par- ticular, suggest that the metal-alkane bond in s-complexes is formed by donation of electron 1Department of Physics and Astronomy, Uppsala University, 751 20 Uppsala, Sweden. 2Center for Free-Electron Laser Science, Department of Physics, University of Hamburg, 22761 Hamburg, Germany. 3Department of Physics, Stockholm University, 106 91 Stockholm, Sweden. 4Paul-Scherrer Institute, CH-5232 Villigen PSI, Switzerland. 5Institute of Nuclear Physics, Polish Academy of Sciences, PL-31342 Krakow, Poland. 6European XFEL GmbH, 22869 Schenefeld, Germany. *Corresponding author. Email: raphael.jay@physics.uu.se (R.M.J.); ambar.banerjee@physics.uu.se (A.B.); philippe.wernet@physics.uu.se (P.W.) †These authors contributed equally to this work. ‡Present address: MCA Engineering GmbH, 80807 Munich, Germany. §Present address: Proximion AB, 164 40 Kista, Sweden. ¶Present address: Department of Physics and Astronomy, Uppsala University, 751 20 Uppsala, Sweden. #Present address: Stanford PULSE Institute, SLAC National Accelerator Laboratory, Stanford University, Menlo Park, CA 94025, USA. **Present address: Physical Sciences Division, Pacific Northwest National Laboratory, Richland, WA 99352, USA. Jay et al., Science 380, 955–960 (2023) 2 May 2023 Table 1. Mulliken charge and orbital properties of the CpRh(CO)-octane and Rh(acac)(CO)- octane s-complexes (B3LYP level of theory). s-complex Rh Mulliken charge LUMO character Rh 4d (%) Cp or acac (%) CO (%) Octane (%) CpRh(CO)-octane ..................................................................................................................................................................................................................... Rh(acac)(CO)-octane ..................................................................................................................................................................................................................... 0.37 0.46 28.8 15.2 39.3 52.2 4.2 1.5 6.2 9.2 1 of 6 RES EARCH | R E S E A R C H A R T I C L E of main-edge features (Fig. 1D). The temporal evolution of the pre-edge peak intensity, shown as a time trace in Fig. 1E, is well described by a biexponential decay to a metastable species (reduced c2 = 1.11; see supplementary mate- rials for a kinetic model). The two time con- stants are assigned to CO dissociation from excited states of CpRh(CO)2 within 370 ± 50 fs followed by octane association within 2.0 ± 0.1 ps. These assignments agree with the time- scales for ligand substitution in other metal carbonyls from previous femtosecond mea- surements (28, 34). Our experiment establishes the timescale of formation of the CpRh(CO)- octane s-complex, and the spectrum at 10 ps in Fig. 1D constitutes a direct fingerprint of how metal-ligand charge-transfer interactions change upon substituting a CO with an alkane ligand. On nanosecond timescales, the disap- pearance of the s-complex pre-edge peak (time trace at 3004.4 eV in Fig. 1F) and the simul- taneous emergence of a positive absorption feature (time trace at 3006.6 eV in Fig. 1F and transient spectrum at nanosecond delay times in Fig. 1D) reflect how the metal-ligand charge-transfer interactions further change upon C–H activation by oxidative addition. Fig. 1. Mechanistic model and time-resolved XAS of C–H activation by CpRh(CO)2 in octane solution. (A) Schematic of C–H activation by CpRh(CO)2 via photoextrusion of CO followed by alkane complexation and oxidative addition. hn, UV photon. (B) Orbital-specific metal-ligand charge- transfer interactions for metal-alkane and metal-carbonyl bonds. (C) Schematic of the experiment with UV–laser pump pulses triggering the reaction and x-ray pulses probing orbital evolution as a function of time delay between pump and probe pulses. Reaction intermediates and products [as well as ground-state CpRh(CO)2] are characterized by detecting the Rh fluorescence as a measure of the Rh-specific x-ray absorption (see supplementary materials). (D) Steady-state and transient Rh L3-edge absorption spectra at indicated pump-probe time delays as well as a schematic depiction of the L-edge absorption process [difference spectra are plotted relative to the edge-jump of the steady-state spectrum (intensity at 3015 eV), which is normalized to 1; steady-state and difference spectrum at delays >190 ns are scaled for illustration]. Rel. abs., relative absorption. (E and F) Time traces (intensities versus time delay) measured at indicated x-ray photon energies with (E) femtosecond and (F) picosecond time resolution. In (E), the gray, orange, and purple shaded regions represent the relative populations of the CpRh(CO)2 excited state, the CpRh(CO) fragment, and the CpRh(CO)-octane s-complex, respectively. Jay et al., Science 380, 955–960 (2023) 2 May 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Both time traces are modeled with a single exponential (reduced c2 = 1.01), yielding a time constant of 14 ± 2 ns, which is in ex- cellent agreement with the ~14 ns for C–H activation of octane with CpRh(CO)2 from time-resolved IR measurements (18). Comparison with theory This assignment of the transient x-ray absorp- tion spectra is further validated and detailed by the calculated spectra in Fig. 2A. Because shapes and intensities of the measured spectra are well reproduced, we can robustly assign x-ray transitions to underlying charge-transfer interactions (see supplementary materials for computational details and discussion of devia- tions between experiment and theory). We use the experiment-theory comparison to extract the orbital correlation diagram shown in Fig. 2B. Fig. 2. X-ray absorption signatures of s-formation and C–H activation by oxidative addition. (A) Experi- mental spectra at time t = 10 ps and >190 ns (top) compared with calculated spectra of CpRh(CO)- octane and CpRh(CO)-H-R [middle, calculated on the B3LYP level of theory (43)]. L3-edge transitions and spectra calculated for intermediate structures (bottom) illustrate the interconversion of spectral features from reactant to product along the C–H activation reaction coordinate (39). Calculated difference spectra are scaled such that the CpRh(CO)- octane difference spectrum matches the pre-edge intensity of the experi- mental spectrum at 10 ps. Vertical lines indicate positions of spectral fingerprints a, b, and c. (B) Correla- tion diagram between the valence orbitals of CpRh(CO)2, CpRh(CO)- octane, and CpRh(CO)-H-R detailing the interconversion of orbital energies and character upon ligand substitution and C–H activation. The calculated orbital plots represent the antibonding counterpart of the bonding interactions that are sche- matically shown in Fig. 1B. For illustration, calculated orbitals are displayed with varying isovalues (see supplementary materials). (C) Calculated free energies (top), Rh 4d character of LUMO+1 and LUMO+3 orbitals (middle), and oscillator strengths of transitions into LUMO+1 and LUMO+3 orbitals (bottom) as a function of reaction coordinate of oxidative addition. arb. u., arbitrary units. Jay et al., Science 380, 955–960 (2023) 2 May 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Importantly, this correlation diagram, which is based on robust experimental observations, relates orbital interactions from ligand substi- tution and s-complex formation to C–H bond breaking and oxidative addition. Two major effects in metal-ligand bonding of the CpRh(CO)-octane s-complex compared with CpRh(CO)2 are reflected in the 10-ps transient spectrum. First, substituting the strong-field CO ligand with the weakly inter- acting octane stabilizes (decreases the ener- gy of) the Rh 4d-derived LUMO orbital (Fig. 2B). This is directly reflected in a decrease of 2p→LUMO transition energies: The pre-edge peak that is due to 2p→LUMO transitions is shifted to lower energy in the s-complex (3004.2 eV) compared with CpRh(CO)2 (3006 eV; Fig. 2A). Second, an overall reduced degree of back-donation in the s-complex compared with CpRh(CO)2 lowers the hybridization of ligand orbitals with Rh 4d orbitals. This dimi- nishes intensities of the Rh 2p transitions to ligand-derived orbitals in the s-complex and causes the depletion in the main-edge region of 3006 to 3009 eV (Fig. 2A). For the subsequent C–H bond breaking and oxidative addition step from the s-complex to the metal alkyl hydride, calculations of the free-energy landscape shown in the top panel of Fig. 2C suggest a barrier of ~7 kcal/mol and an exothermic reaction. The underlying reaction coordinate is constructed from a Nudged-Elastic-Band (NEB)/TPPSh/Def2-TZVP computation (35–37). Using the geometries of this reaction path scan, the free-energy land- scape was computed at the DLPNO-CCSD(T)/ Def2-TZVP level of theory (38). Although we do not experimentally observe the intermediate structures along the reaction coordinate, the L3-edge x-ray absorption spectra computed for these structures relate the spectral changes from the s-complex reactant to the metal alkyl hydride product, which we observe. The key spectral fingerprint regions (denoted as a, b, and c in Fig. 2A) can be assigned to excitations of Rh 2p electrons predominantly into the LUMO, LUMO+1, LUMO+2, and LUMO+3 orbitals (LUMO+4, +5, … are not discussed because they contribute to a negligible degree only). Changes of the features a to c hence report on the combined transformations of the four lowest unoccupied orbitals upon C–H bond breaking and oxidative addition. As detailed in the following paragraphs, feature a re- flects changes in metal-alkane orbital over- lap as metal-alkane bond distances change, feature b reports on the oxidation of the metal, and feature c reflects changes in metal-ligand back-donation. In line with previous work (39), our cal- culated reaction coordinate describes the C–H bond moving toward the Rh center and, at the same time, the C–H bond elongating and breaking until the individual Rh–C and Rh–H Jay et al., Science 380, 955–960 (2023) 2 May 2023 Fig. 3. X-ray orbital view of reactivity modulations in C–H activation by varying the ligand environ- ments in s-complexes. (A) Schematic of the LUMO orbitals of CpRh(CO)-octane and Rh(acac)(CO)-octane with variations in metal-ligand bonding (charge-transfer indicated by arrows) and their effect on the affinity for oxidative addition (calculated free energies). Me, methyl. (B) Calculated difference spectra of CpRh(CO)-octane and Rh(acac)(CO)-octane compared with transient difference L3-edge absorption spectra measured at SwissFEL at a pump-probe delay time of 10 ps. For comparison, the experimental Rh(acac)(CO)- octane spectrum is scaled to match the depletion of the CpRh(CO)-octane. This scaling is validated by the excellent agreement with the calculated spectra, which are shown with the same scaling as in Fig. 2A (see supplementary materials). bonds are established (see schematic of the reaction coordinate in Fig. 2A, bottom). As a consequence of these atomic rearrangements along the reaction coordinate, the Rh 4d-derived LUMO orbital shifts to higher energies be- cause of increasing orbital overlap with the approaching C–H group with minor changes in hybridization (Fig. 2B and fig. S7). The cor- responding increase of 2p→LUMO transition energies is directly observed experimentally by the disappearance of the pre-edge feature a in the spectrum upon transformation of the s-complex to the alkyl hydride product: The 2p→LUMO transitions shift to higher energy and merge with the main edge of the spec- trum, thereby contributing to the generation of feature b in the spectrum of the CpRh(CO)– H–R alkyl hydride product. LUMO+1 in the s-complex is the energeti- cally lowest ligand-derived orbital with domi- nant CO p* character and with some Rh 4d admixture due to Rh-CO back-donation (see orbital plots in Fig. 2B). Upon oxidative ad- dition, LUMO+1 shifts to slightly lower energy and, importantly, gains considerable Rh 4d character (see calculated Rh 4d character in Fig. 2C). The increase is so substantial that the Rh 4d character becomes the dominating contribution. This can be interpreted as an effective transformation of the former ligand- derived orbital into a second unoccupied Rh 4d-derived orbital (in addition to the LUMO; see orbital plots in Fig. 2B). The increase of Rh 4d character directly scales with an increase of oscillator strength of the Rh 2p→LUMO+1 transitions (Fig. 2C). Feature b hence emerges as a strong peak, drawing intensity from both the LUMO+1 transforming into a second un- occupied Rh 4d-derived orbital and the LUMO shifting to higher energy and merging with the main edge (Fig. 2A). The emergence of feature b thus reflects the combined electronic-structure effects of C–H bond cleavage (LUMO destabi- lization) and oxidative addition (LUMO+1 transformation). In particular, the oxidation of the metal center from a Rh(I) (d8) to a Rh(III) (d6) configuration is evidenced by the emer- gence of two unoccupied Rh 4d orbitals. The increase of the Rh oxidation state sub- stantially destabilizes LUMO+2 (Fig. 2B) and slightly reduces its Rh 4d character (see fig. S7). As the second CO p* orbital, its destab- ilization thus directly reflects a decrease in back-donation from the oxidized metal onto CO p*. This effect has also been associated with the shift of CO marker modes to higher energy upon C–H activation (17), consistent with our results. Finally, LUMO+3 in the s-complex constitutes the octane C–H s* orbi- tal, which exhibits weak Rh 4d admixture be- cause of low back-donation from Rh to C–H (orbital plot in Fig. 2B). Back-donation, how- ever, increases as the C–H bond is broken and the covalent Rh–C and Rh–H bonds are formed, as evidenced by the substantial increase in Rh 4d character in LUMO+3 (see Rh 4d character in Fig. 2C and orbital plots in Fig. 2B). Ac- cordingly, the oscillator strengths of Rh 2p→LUMO+3 transitions also strongly increase upon oxidative addition (Fig. 2C). Together with the transitions into LUMO+2 shifting toward higher energies, this causes the for- mation of the strong peak c in the alkyl hydride spectrum, which exhibits an intensity similar 4 of 6 RES EARCH | R E S E A R C H A R T I C L E to the steady-state spectrum of CpRh(CO)2 and one considerably stronger than that in the s-complex (Fig. 2A). Experimentally, this is reflected in negligible intensities in the alkyl hydride difference spectrum at the energies of feature c compared with the strong bleaching in the transient s-complex spectrum. A more stable s-complex By experimentally observing individual charge- transfer interactions, we verify the validity of orbital correlation diagrams along C–H activation reactions that were previously derived from quantum chemical calculations alone (40). Our approach allows us, in particular, to ex- pand upon established notions of charge- transfer interactions between the metal and the C–H group by experimentally assessing the critical role of additional orbital interac- tions between the metal and the spectator ligands. We further demonstrate this here by evaluating how different s-complexes exhibit different reactivities toward C–H activation owing to specific differences in orbital inter- actions as a result of their different ligand environments. It has previously been shown that replacing the Cp moiety with an acetyl- acetonate (acac) group leads to a stable s-complex, which, however, does not proceed to oxidative addition of the C–H bond (41). Our calculations shown in Fig. 3A suggest a 4.2 kcal/mol stabilization of the Rh(acac) (CO)-octane with respect to the CpRhCO- octane s-complex. Together with the endo- thermic free-energy profile we calculated, this renders the C–H activated product un- favorable. We find the extra stabilization of Rh(acac)(CO)-octane to be predominantly due to a higher donation from the octane onto the Rh center. This stronger donation is favored by the higher charge deficiency at the Rh in the case of the more ionic bond between Rh and the acac group compared with the bond between Rh and Cp (see the Mulliken charges in Table 1). Our calculations predict this variation in ionicity and the related variation in reactivity for C–H activation to manifest in the x-ray absorp- tion difference spectra of the two s-complexes as shown in Fig. 3B. Our experiment directly confirms this prediction. In quantitative agree- ment with theory, the measured spectrum of Rh(acac)(CO)-octane shows a higher pre-edge intensity than CpRh(CO)-octane (Fig. 3B). We find this difference to be due to a higher Rh 4d character in the LUMO (at the expense of a lower hybridization with the acac group; see Table 1), which causes the more intense 2p→LUMO pre-edge transitions in Rh(acac) (CO)-octane. This higher Rh 4d character, which directly correlates with higher Rh ionicity, renders the reaction step to the Rh(acac)(CO)-H-R species endothermic. A more charge-deficient Rh(I) center—in the Jay et al., Science 380, 955–960 (2023) 2 May 2023 case of acac, having a lower propensity to be further oxidized to Rh(III)—is consistent with and extends established trends in alkane oxi- dative addition (7). We thus establish a direct measure of how the lower hybridization of Rh 4d with spectator-ligand orbitals in the more ionic bond to the acac ligands modulates reactivity for C–H activation by unfavorably changing the balance of charge-transfer inter- actions that bind (alkane-to-metal s-donation) versus those that break the C–H bond (pro- pensity for oxidation via metal-to-alkane back-donation). Our results demonstrate the value of time- resolved, metal-specific, L-edge x-ray absorp- tion spectroscopy for understanding, on an orbital level, which factors determine reactivity for C–H activation at a metal complex. We anticipate that our approach will be used in the future to systematically screen s-complexes and alkyl hydride reaction products to provide a distribution of valence orbital energies and character as measures of metal-alkane bond stability and propensity toward C–H activation with oxidative addition and, potentially, other mechanisms (40). With C–H activation ranging from nucleophilic to electrophilic, depending on the relative weight of charge donation and back-donation, the here established experi- mental observables can be used to ascertain where in the range of mechanisms a probed system lies. Such insight can then be used to pin the results from computational studies that correlate valence electronic structure with reactivity. We envision this approach to extend established trends for reactivity (7) by pro- viding experimentally verified correlations between metal-ligand charge-transfer inter- actions and reactivity for orbital-level control of C–H activation. RE FERENCES AND NOTES 1. R. G. Bergman, Nature 446, 391–393 (2007). J. A. Labinger, J. E. Bercaw, Nature 417, 507–514 (2002). 2. 3. B. A. Arndtsen, R. G. Bergman, T. A. Mobley, T. H. Peterson, Acc. Chem. Res. 28, 154–162 (1995). 4. K. I. Goldberg, A. S. Goldman, Acc. Chem. Res. 50, 620–626 5. (2017). J. K. Hoyano, A. D. McMaster, W. A. G. Graham, J. Am. Chem. Soc. 105, 7190–7191 (1983). 6. A. H. Janowicz, R. G. Bergman, J. Am. Chem. Soc. 104, 7. 352–354 (1982). J. Hartwig, Organotransition Metal Chemistry: From Bonding to Catalysis (University Science Books, 2010). 8. C. Hall, R. N. Perutz, Chem. Rev. 96, 3125–3146 (1996). 9. W. D. Jones, Acc. Chem. Res. 36, 140–146 (2003). 10. J. B. Asbury, H. N. Ghosh, J. S. Yeston, R. G. Bergman, T. Lian, Organometallics 17, 3417–3419 (1998). 11. S. A. Bartlett et al., J. Am. Chem. Soc. 141, 11471–11480 (2019). 12. D. J. Lawes, S. Geftakis, G. E. Ball, J. Am. Chem. Soc. 127, 4134–4135 (2005). 13. F. M. Chadwick et al., J. Am. Chem. Soc. 138, 13369–13378 (2016). 14. S. Geftakis, G. E. Ball, J. Am. Chem. Soc. 120, 9953–9954 (1998). 15. J. D. Watson, L. D. Field, G. E. Ball, Nat. Chem. 14, 801–804 (2022). 16. E. P. Wasserman, C. B. Moore, R. G. Bergman, Science 255, 315–318 (1992). 17. S. E. Bromberg et al., Science 278, 260–263 (1997). 18. M. W. George et al., Proc. Natl. Acad. Sci. U.S.A. 107, 20178–20183 (2010). 19. A. L. Pitts et al., J. Am. Chem. Soc. 136, 8614–8625 (2014). 20. M. C. Asplund et al., J. Am. Chem. Soc. 124, 10605–10612 (2002). 21. E. A. Cobar, R. Z. Khaliullin, R. G. Bergman, M. Head-Gordon, Proc. Natl. Acad. Sci. U.S.A. 104, 6963–6968 (2007). 22. J. Y. Saillard, R. Hoffmann, J. Am. Chem. Soc. 106, 2006–2026 (1984). 23. P. E. M. Siegbahn, M. Svensson, J. Am. Chem. Soc. 116, 10124–10128 (1994). 24. D. Balcells, E. Clot, O. Eisenstein, Chem. Rev. 110, 749–823 (2010). 25. R. M. Jay, K. Kunnus, P. Wernet, K. J. Gaffney, Annu. Rev. Phys. Chem. 73, 187–208 (2022). 26. Y. Kim et al., J. Phys. Chem. Lett. 12, 12165–12172 (2021). 27. B. E. Van Kuiken et al., J. Phys. Chem. Lett. 3, 1695–1700 (2012). 28. P. Wernet et al., Nature 520, 78–81 (2015). 29. W. Gawelda et al., J. Am. Chem. Soc. 128, 5001–5009 (2006). 30. A. A. Cordones et al., Nat. Commun. 9, 1989 (2018). 31. B. E. Van Kuiken et al., J. Phys. Chem. A 117, 4444–4454 (2013). 32. F. De Groot, Coord. Chem. Rev. 249, 31–63 (2005). 33. R. K. Hocking et al., J. Am. Chem. Soc. 128, 10442–10451 (2006). 34. A. G. Joly, K. A. Nelson, Chem. Phys. 152, 69–82 (1991). 35. V. Ásgeirsson et al., J. Chem. Theory Comput. 17, 4929–4945 (2021). 36. J. Tao, J. P. Perdew, V. N. Staroverov, G. E. Scuseria, Phys. Rev. Lett. 91, 146401 (2003). 37. F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 7, 3297–3305 (2005). 38. Y. Guo et al., J. Chem. Phys. 148, 011101 (2018). 39. R. H. Crabtree, E. M. Holt, M. Lavin, S. M. Morehouse, Inorg. Chem. 24, 1986–1992 (1985). 40. D. H. Ess, W. A. GoddardIII, R. A. Periana, Organometallics 29, 6459–6472 (2010). 41. T. P. Dougherty, W. T. Grubbs, E. J. Heilweil, J. Phys. Chem. 98, 9396–9399 (1994). 42. R. M. Jay et al., Data for “Tracking C–H activation with orbital resolution.” Zenodo (2023); https://doi.org/10.5281/zenodo. 7837518. 43. A. D. Becke, J. Chem. Phys. 98, 1372–1377 (1993). AC KNOWLED GME NTS We acknowledge the Paul Scherrer Institut, Villigen, Switzerland, for provision of beamtime at the Alvra beamline of SwissFEL as well as at the PHOENIX beamline of the Swiss Light Source (SLS). We thank R. Wetter and C. Frieh for their excellent technical support. The computations were partly enabled by resources provided by the Swedish National Infrastructure for Computing (SNIC) at UPPMAX, which is partially funded by the Swedish Research Council through grant agreement nos. 2021-22968 and 2022-22975. The computations were also partly enabled by resources provided by the National Academic Infrastructure for Supercomputing in Sweden (NAISS) and the SNIC at the National Supercomputer Centre in Sweden (NSC) and the PDC (Parallelldatorcentrum) Center for High Performance Computing, which are partially funded by the Swedish Research Council through grant agreement nos. 2022-06725 and 2018-05973. Funding: A.B. and P.W. acknowledge funding from the Carl Tryggers Foundation (contract CTS 19: 399). P.W. acknowledges funding from the Swedish Research Council (grant agreement no. 2019-04796). J.H. and N.H. acknowledge funding from the Cluster of Excellence “CUI: Advanced Imaging of Matter” of the Deutsche Forschungsgemeinschaft (DFG), EXC 2056, project ID 390715994. R.-P.W. acknowledges funding from the German Ministry of Education and Research (BMBF), project ID 05K19GU2. V.K., A.K., and C.B. acknowledge support from the Swiss National Science Foundation (SNSF) through the NCCR:MUST. A.W. acknowledges the National Science Centre, Poland (NCN), for partial support through grant no. 2019/03/X/ST3/00035. Author contributions: R.M.J., A.B., and P.W. originated the project concept. R.M.J., P.J.M.J., C.C., C.B., C.M., N.H., G.S., T.H., and P.W. planned and conceived the experiments. R.M.J., T.L., R.S., R.-P.W., J.H., E.V.B., V.K., A.K., A.W., D.O., C.A., P.J.M.J., C.N.B., C.C., C.B., G.S., T.H., and P.W. executed the experiments. R.M.J., T.L., H.W., C.C., and P.W. analyzed the experimental data. R.M.J., A.B., M.R.C., and M.O. performed the theoretical calculations. R.M.J., A.B., and P.W. wrote the paper 5 of 6 RES EARCH | R E S E A R C H A R T I C L E with input from all the authors. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data presented in the main text and the supplementary materials are freely available through Zenodo (42). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf8042 Materials and Methods Supplementary Text Figs. S1 to S11 Table S1 References (44–65) Movies S1 to S4 Submitted 20 January 2023; accepted 2 May 2023 10.1126/science.adf8042 Jay et al., Science 380, 955–960 (2023) 2 May 2023 6 of 6
10.1126_science.adf7553
a coherent superposition of two or more states with well-separated phase-space distributions. Experimental setup Our device, which we call a ℏBAR, as in pre- vious works (19), consists of a high-overtone bulk acoustic-wave resonator (HBAR) coupled to a superconducting transmon qubit. The transmon qubit allows us to create, control, and read out phonon states in the HBAR. Qubit and HBAR are fabricated on separate sapphire chips, which are subsequently flip chip bonded into the final device (Fig. 1A). A dome of piezoelectric aluminum nitride on the HBAR chip coherently couples the electric field of the qubit with the strain field of the resonator modes. The device is placed inside a three-dimensional Al cavity, which allows us to control the qubit with the standard circuit quantum electrodynamics toolbox (20). The acoustic free spectral range is ∼12 MHz, and frequency tuning the qubit via a microwave drive–induced Stark shift allows us to address several longitudinal phononic modes. The acoustic lattice oscillations are localized within a Gaussian mode with waist w0 ¼ 27 mm and length L ¼ 435 mm, giving a mode volume of 0L ≈ 0:001 mm3 [see supplementary text pw2 (21) section B for details]. More details about this circuit quantum acoustodynamics (cQAD) system (19, 22) and the device (23) can be found in previous works. In the classical picture, one can imagine a coherent state jai in the phonon mode as a coherent displacement of the atomic lattice with an amplitude proportional to a. In the quantum picture, an example of a cat state is a quantum superposition of two coherent states with opposite displacement amplitudes, leading to the physical interpretation of such a state as the superposition of two oscillations RES EARCH QUANTUM MECHANICS Schrödinger cat states of a 16-microgram mechanical oscillator Marius Bild1,2†, Matteo Fadel1,2†*, Yu Yang1,2†, Uwe von Lüpke1,2, Phillip Martin1,2, Alessandro Bruno1,2, Yiwen Chu1,2* According to quantum mechanics, a physical system can be in any linear superposition of its possible states. Although the validity of this principle is routinely validated for microscopic systems, it is still unclear why we do not observe macroscopic objects to be in superpositions of states that can be distinguished by some classical property. Here we demonstrate the preparation of a mechanical resonator in Schrödinger cat states of motion, where the ~1017 constituent atoms are in a superposition of two opposite-phase oscillations. We control the size and phase of the superpositions and investigate their decoherence dynamics. Our results offer the possibility of exploring the boundary between the quantum and classical worlds and may find applications in continuous-variable quantum information processing and metrology with mechanical resonators. Q uantum mechanics is one of the most successful scientific theories ever for- mulated. However, from the early days of quantum mechanics until now, it has been unclear why quantum phenomena, such as state superpositions, are never ob- served in the macroscopic world. In his 1935 work (1), Erwin Schrödinger imagined a de- vice able to poison a cat as a consequence of a radioactive decay, concluding that the super- position of an atom being “decayed” and “not decayed” could be mapped onto a superposi- tion of the cat being simultaneously “dead” and “alive.” There are two aspects of this hy- pothetical scenario that make it seem absurd and counterintuitive: First, a cat is a macro- scopic, everyday object; and second, “dead” and “alive” are states that are mutually exclusive within our classical experience. Many explanations have been proposed as to why we may never encounter a cat in such an unfortunate situation. Macroscopic objects may simply be too complex and subject to too many sources of decoherence to sustain a super- position of classically distinct states. Other theories introduce additional effects beyond standard quantum mechanics, such as wave function collapse due to intrinsic stochastic noise or gravitational decoherence (2). These effects are typically expected to scale with the mass of the system and the distinctness of the states that are superposed. Therefore, observ- ing state superpositions in massive objects is of key importance for exploring the validity range of quantum mechanics as we know it. Beyond its fundamental interest, preparing and detecting Schrödinger’s cat states is essential 1Department of Physics, ETH Zürich, 8093 Zürich, Switzerland. 2Quantum Center, ETH Zürich, 8093 Zürich, Switzerland. *Corresponding author. Email: fadelm@phys.ethz.ch (M.F.); yiwen.chu@phys.ethz.ch (Y.C.) †These authors contributed equally to this work. for applications in quantum technologies. Main examples include Heisenberg-limited parame- ter estimation protocols (3, 4) and error-protected quantum information processing (5, 6). There have been many experimental dem- onstrations of Schrödinger cat states (which we will call “cat states” from here on). These include superpositions of internal and mo- tional degrees of freedom in trapped ions (7, 8), phase-space superpositions of electromagnetic waves in both the optical (9, 10) and microwave domains (11–13), Greenberger–Horne–Zeilinger states (14, 15), current superpositions in super- conducting quantum interference devices (16), and spatial superpositions of atomic clouds (17) and large molecules (18). We experimen- tally demonstrate the preparation of cat states in the motional degree of freedom of a solid- state mechanical resonator. Given the variety of definitions found in previous works, we define a cat state of a harmonic oscillator as Fig. 1. Illustration of the h– BAR device and system evolution. (A) Schematics of the ħBAR device. The HBAR chip (top) has a layer of piezo- electric aluminum nitride (orange) and supports standing acoustic waves (pink). The transmon qubit on the lower chip has a circular antenna to couple with the HBAR. The inset shows the superposition of two opposite-phase oscillations of atoms in the crystal lattice. (B) Simulated evolution without decoherence of the qubit jei state population Pjei and purity g under the JC interaction when the qubit is initialized in j(cid:2)Zi and the phonon in a coherent state. (C) Illustration of the evolution of an initial phonon coherent state (red circle on the left) in phase space. The blue (yellow) crescent shapes indicate the state jFþi (jF(cid:2)i), which is the phonon state when the qubit is initialized in j þXi (j(cid:2)Xi). Interference fringes appear around time tC when the qubit is prepared in a superposition of jþXi and j(cid:2)Xi. Around the revival time tR, the two phonon states again overlap (purple). Bild et al., Science 380, 274–278 (2023) 21 April 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E of the atomic lattice with the same frequency wp and relative phase p. Considering a snap- shot in time where both oscillations are at their displacement maximum, Schrödinger’s cat being in a superposition of dead and alive is analogous to a superposition of atoms in the HBAR being in two distinct positions in space, as illustrated in the inset of Fig. 1A. Here, we define the positions as distinct when their separation is larger than the fluctua- tions due to quantum, thermal, or other sources of noise. Generating cat states To realize a cat state in our system, we use the Jaynes-Cummings (JC) interaction with the qubit and phonon on resonance (11, 24). The in- teraction Hamiltonian is (cid:1) H=ℏ ¼ g0 sþa þ s(cid:2)a† (cid:3) ð1Þ (cid:1) p ffiffiffi n where g0 is the coupling strength between qubit and phonon mode, sþ is the raising op- erator for the qubit, and a† is the raising op- erator for the phonon mode. This results in Rabi oscillations between the states je; n (cid:2) 1i p ffiffiffi , where jgi (jei) is the and jg; ni at a rate g0 n qubit ground (excited) state andjni is the Fock state of n phonons. As a consequence of this p scaling, if the phonon mode is prepared in a coherent state with large enough amplitude and the qubit is prepared in jgi or jei, their coherent interaction rapidly dephases. Hence, the oscillations of the qubit population “col- lapse” (Fig. 1B) with a decaying amplitude (cid:3) ð proportional to (25, 26) exp (cid:2) t=tcollapse , where tcollapse ¼ =g0 is the collapse time in the limit of a ≫ 1. At this time, the qubit and phonon states are entangled. This can be seen in Fig. 1B as a minimum in the qubit state pu- rity g tð Þ ¼ Tr rq around tcollapse, where rq tð Þ is the reduced density matrix of the qubit. Notably, owing to the quantized phonon energy and the consequent discrete oscilla- tion frequency spectrum, the oscillations revive in finite time (24). For a ≫ 1, this revival occurs at tR ¼ 2pa=g0 (27). Between the collapse and revival, at time tR=2, the qubit and phonon disentangle from each other. The state being separable at tR=2 coincides with the occur- rence of a superposition of two distinct states in phase space, realizing a cat state in the pho- non mode (11, 21, 24, 28). ðtÞ2 ffiffiffi 2 Þ2 (cid:3) (cid:1) A more intuitive explanation for the origin of the cat state comes from the time evolution of the reduced phonon state in phase space. As illustrated in Fig. 1C and shown in sup- plementary text section A, if the qubit is ini- tialized in the state TXj and in the limit of large a, the evolution leads to a rotation in phase space with an angular veloc- ity ∓ g0=2a j and a distortion of the coherent j states. We call the resulting statesjFT tð Þi, whose full expressions are given in the supplemen- i ≡ ej i T gj i ffiffiffi 2 Þ= p ð Fig. 2. Collapse and revival dynamics. (A) Experimental sequence for observing collapse and revivals dynamics and for preparing cat states (details in the main text). (B) Measured qubit population and state purity. The solid and dashed black lines are the simulation results of the qubit population and purity, respectively. Three time points of particular interest are highlighted (dashed lines): initial state time, cat state time (tC), and revival state time (tR). (C) Measured Wigner function of the phonon state at the three time points. Axes are the real and imaginary parts of the complex displacement amplitude b used during Wigner tomography (23). The black crosses indicate the positions of the two coherent states composing the fitted CSS state Eq. 2) (D) Corresponding simulated Wigner functions. ð p j(cid:2)X i tary text (eq. S15). If the the qubit is initialized ffiffiffi in the state jTZi ≡ þX i T Þ= j 2 , the phonon state will evolve into Fþ tð Þi T jF(cid:2) tð Þi j as shown in Fig. 1C. At time tR=2, the two state components jFTi have covered a rotation angle of ∓p=2 around a circle of radius a, maximiz- ing their separation in phase space and form- ing a cat state (24). Finally, at the revival time tR , the two phonon state components jFTi have both rotated by a phase of p and approx- imately recombine in phase space. Experimental results We experimentally confirm both the predicted collapse and revival of Rabi oscillations and the creation of mechanical cat states in the phonon mode. The basic sequence used in the experimental demonstration of the JC dynam- ics described above can be seen in Fig. 2A. We displace the phonon mode with a resonant drive of amplitude A to a coherent state with amplitude a. To mitigate any effect of the drive on the qubit state, we then cool the qubit with an ancillary phonon mode (22, 23). The qubit is subsequently prepared in its initial state by applying a drive pulse with variable phase and amplitude. To induce the resonant interaction, we tune the qubit to the phonon mode frequency for a variable interaction time t. Depending on which of the subsystems we want to char- acterize, we choose a measurement sequence that implements the appropriate measure- ment operator. First, we simply measure the qubit excited state population. The resulting data are shown in Fig. 2B for A ¼ 0:35. Here the value for A is a scaling factor for the am- plitude of a microwave drive, which we cali- brate to find a corresponding initial coherent state size of a ¼ 1:75 (supplementary text sec- tion E). As expected, we observe oscillations that collapse after a time tcollapse ≈ 0:9 ms and revive at tR ≈ 6:7 ms. This revival indicates the coherent exchange of energy quanta between the qubit and phonon mode during the resonant interaction. By performing full qubit tomog- raphy after the resonant interaction, we can also reconstruct the reduced density matrix of the qubit subsystem rq and calculate the purity of the qubit state g tð Þ . We confirm a local minimum of g tð Þ around the predicted collapse time tcollapse, followed by a local max- imum around tR=2 (Fig. 2B). We now focus on the time evolution of the phonon subsystem by performing full Wigner tomography of the phonon state after the resonant interaction times t = 0, 2.9 and 7.0 ms. To this end, we use the parity measurement technique established in a previous work (23). To compensate for the effect of qubit dephasing during the parity measurement, we normalize Bild et al., Science 380, 274–278 (2023) 21 April 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Cat state amplitude and phase control. (A) Cat states prepared with different displacement pulse amplitudes A. Top row: Measured Wigner functions; bottom row: analytical state r tCð Þ that best fits the data. The fitted CSS states, with coherent state positions indicated by black crosses, have D = 1.09 (1.43) for A = 0.25 (0.30). (B) Cat state sizes as a function of the displacement amplitude A, obtained from fitting the data in (A) to analytical states r tCð states jCi. Numbers are the fidelity with respect to the fitted state. Error bars show cat state sizes resulting in 1 % deviation in the fidelity (supplementary text section G) (C) Cat states resulting from the initial qubit states indicated by the respective labels. (D) Cross- cuts of interference fringes from (C) for four initial qubit states, obtained by averaging the data between (cid:2)0:11 < Im bð Þ < 0:16. Þ and CSS pected, Fig. 2D shows the results of a master equation simulation of the full experimental protocol with independently measured sys- tem parameters, showing good agreement with the measurements in Fig. 2C. The state we obtained resembles the two- component coherent state superpositions (CSS) jCi ¼ N a1 (cid:1) j i þ eiϑ a2j i (cid:3) ð2Þ (cid:1) (cid:3) ; jCi Ch j a type of cat state that is often invoked in quantum information (5, 6) and parameter estimation protocols (3, 4). Here ja1;2i are two coherent states and N is the appropriate nor- malization constant. We can fit our recon- structed state to Eq. 2 by optimizing a1 , a2, and ϑ for maximum fidelity F rp . Be- cause our state is not centered around the origin in phase space, we use half the phase space dis- tance between the coherent state components D ¼ a1 (cid:2) a2 j=2 as a measure of the cat state j size. This choice is motivated by considering a coherent state superposition centered around the origin in phase space, such that a1 ¼ (cid:2)a2. ffiffiffi , where (cid:1)n is the average (cid:1)n Then D ¼ a1;2 phonon population of the state created. For the state in Fig. 2C, we obtain a state size D ¼ 1:61, corresponding to (cid:1)n ¼ D2 ¼ 2:60, with a fidel- ity of F ≈66%. The smaller state size D com- pared to the initial coherent displacement is a combination of decoherence and the choice of interaction time tC < tR=2, resulting in the two counter-rotating state components not reaching their maximum separation in phase space. The fidelity is lower compared to the fitted analytical state r tCð Þ itself has a finite infidelity to the CSS state jCi. Þ, because r tCð (cid:5) (cid:5) ¼ p (cid:5) (cid:5) all measured parity values to that of the Fock j0i phonon state (supplementary text section D). The measured Wigner functions are shown in Fig. 2C, where axes in phase space are nor- malized by measuring the distribution of pop- ulations in the phonon Fock states for coherent states created with different drive amplitudes (22) (supplementary text section E). From the measured data, we confirm the evolution of the initial coherent state (Fig. 2C, left) into a cat state at tC ¼ 2:9 ms (Fig. 2C, cen- ter), showing two state components clearly distinct in phase space and interference fringes located between them. We choose this value of tC because it corresponds to the measured maximum in the qubit state purity. It de- viates somewhat from the value predicted by using the large a limit, which is tR=2 ≈ 3:3 ms. For the evolution time t ¼ 7:0 ms, the predicted refocusing into a crescent-shaped overlap be- tween the counter-rotating state components can be observed (Fig. 2C, right). To benchmark the cat state and obtain an estimate of its size, we implement a maximum- likelihood reconstruction (29) of the phonon staterp from the measured state with A ¼ 0:35 and tC ¼ 2:9 ms. We then fit the reconstructed state to an analytical expression of the expected phonon state r tCð Þ in the absence of decoher- ence and after tracing out the qubit (supple- mentary text section A). Fixing the interaction time to tC from the experiment, the fit max- Þ by imizes the fidelity between rp and r tCð varying the initial coherent state size afit of r tCð Þ. The result yields a fidelity of F ≈ 76% to an analytical state with initial coherent state size afit ¼ 1:62, which is smaller than the initial displacement a ¼ 1:75 because the expression for r tCð Þ does not include phonon losses. We attribute the infidelity to a com- bination of decoherence and measurement imperfections that lead to additional artifacts in the Wigner function (23). To further con- firm that the phonon state behaves as ex- We can now translate the parameters of the measured cat state into physical properties of the phonon mode, such as the spatial separa- tion between atoms. A state size of D ¼ 1:61 corresponds to a maximal delocalization of 7:0 (cid:3) xZPF, where xZPF is the zero point motion of an equivalent one-dimensional (1D) quan- tum harmonic oscillator (supplementary text section B). Because we are not considering a center-of-mass mode, there is some freedom in choosing xZPF , which is then associated with an effective oscillating mass of the mode. If we choose the root-mean-square (RMS) value of the atomic displacements, we find an effec- ¼ 16:2 mg, corresponding tive mass of M RMS to ∼1017 atoms, delocalized over a distance of 2.1 × 10−18 m (supplementary text section B). In applications such as bosonic encodings of a qubit state, full control over the phase and amplitude of the created cat state is required (5, 6, 30). In the following, we demonstrate this level of control in our experiment. By varying the amplitude A of the phonon displacement drive, we can control the amplitude of the initial coherent state and the size of the resulting cat state. For displacement amplitudes A ¼ 0:25 and 0:30, we create cat states with D ¼ 1:09 eff Bild et al., Science 380, 274–278 (2023) 21 April 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Decoherence of cat states. (A) Measured 1D cuts through the interference fringes of the D ¼ 1:43 cat state for a range of wait times between state creation and measure- ment. (B) Extracted negativities (squares) from each cut versus wait times for three cat state sizes, together with fitted expo- nential decays (solid lines). Both data and fitted curves are normalized to the fitted value at t ¼ 0. (C) Characteristic decay times tcat extracted from the fits in (B) for all three cat state sizes. Error bars are uncertainties extracted from the fit. and 1:43, respectively (Fig. 3A). The fidelities of reconstructions of the measured states to Þ are given in Fig. both a CSS state and r tCð 3B. The best fit r tCð Þ is plotted in the lower row of Fig. 3A, showing good qualitative agree- ment with the data. As before, the finite fidel- ities and lower-contrast fringes of the measured Þ arise states as compared to the best-fit r tCð mainly from decoherence of the state during measurement. In the two-component cat state encoding of a qubit, the six cardinal points of the Bloch sphere are given by two coherent states ja1i and ja2i, along with their four superpositions with p=2 difference in the phase ϑ (Eq. 2). We can prepare similar states by initializing the transmon qubit state in all six cardinal points jTY i; jTZi of its Bloch sphere before per- TX i; j forming the cat state generation protocol. The preparation of jTX i and jTY i is calibrated using collapse and revival measurements as a function of the qubit drive phase (supplemen- tary text section A). The results for A ¼ 0:35 and tC ¼ 2:10 ms are shown in Fig. 3C. We ob- serve a distorted coherent state located on the upper (lower) half of phase space for the qubit initially in jTX i. This separation in phase space is expected from the opposite rotation directions between the phonon states when the qubit is prepared in jTX i (Fig. 1C). The TY i; jTZi states then give rise to four cat states that differ in phase by p=2, as can be observed in the phases of the interference fringes in Fig. 3 D. The ini- tial energy of the qubit, and thus of the total system, is not the same for all six scenarios, resulting in slightly different sizes for the cat states. In the limit of large cat state size, this difference becomes negligible, and the pho- non subspace maps onto that of the cat state encoding. j Superposition states are nonclassical states that are notoriously prone to decoherence. We now investigate the quantum-to-classical tran- sition of different-sized cat states by letting them evolve freely for a varying wait time t before performing Wigner tomography. Dur- ing this evolution, the qubit is far detuned from the phonon mode. In particular, we fo- cus on a slice through the Wigner function’s interference fringes at Im(b) = 0, which high- lights the nonclassical features of the super- position. Figure 4A shows the time evolution of this slice for the D ¼ 1:43 cat state. We ob- serve that the negative features disappear on a time scale much faster than T ph ≈ 84 ms, the 1 energy relaxation time of the phonon mode. ð Þ ð Þ ð j(cid:2)W b; t As a measure for the nonclassicality of the state, we extract the-time dependent negativity (31), defined as d tð Þ ≡ ∫ W b; t Þdb. ð j Here,W b; t Þ is the measured Wigner function of the cat state at time t, and the integration is over the 1D slice in phase space parameter- ized by the complex displacement amplitude b. Figure 4B shows the resulting d tð Þ for the three different cat state sizes of Fig. 3B. We fit each dataset to an exponential decay plus a constant offset. The offset in the measured Wigner values arises from the fact that our Wigner tomography is not performed in the ideal dispersive limit (23). The extracted decay time scales tcat are plotted in Fig. 4C. We show in section H of the supplementary text that, in j, d tð Þ decays exponentially the limit of large aj (cid:3) (cid:1) with a time constant tcat ¼ T ph = 2 aj . How- 1 j, tcat deviates from this ex- ever, for small aj pression and is in fact dependent on properties of the exact state, such as the phase of the super- position. The data in Fig. 4C show the ex- pected qualitative behavior of faster-decaying negativity for larger-sized cat states, and we present a more detailed quantitative analysis in the supplementary text (21). j2 Concluding remarks Our results show the generation of cat states in a microgram-mass solid-state mechanical mode using the tools of cQAD and pave the way toward using such systems for tests of wave function collapse models (32). These tests would benefit from larger-sized cat states, resonators with higher masses (33), and longer phonon lifetimes. To facilitate comparison with other mechanical resonators and theoretical models that often consider center-of-mass motion, we note that the HBAR mode is a standing wave in which a half-wavelength section approxi- mates a center-of-mass mode where all atoms oscillate in the same direction. Such a section has a mass on the order of 30 ng, obtained by dividing the total effective mass of 16 mg by the longitudinal mode number of ∼500. The maximum size of the cat state that we can prepare is currently limited by our device parameters, including both the qubit and pho- non decoherence rates. The latter is especially important given that, in general, the decoher- ence rate of the cat state is proportional to the square of the cat state size D . Furthermore, additional improvements to the properties of qubit and phonon resonator would enable al- ternative cat state generation protocols that can in principle lead to states with a higher fidelity to, for example, a CSS state (12, 13). We point out, however, that although CSS states represent a useful benchmark (because they have been extensively studied for applications such as quantum information and quantum metrology), many of their salient features are present already in the states that we have demonstrated. These include the phase-space separation of state components, which is im- portant for error protection of encoded qubits (30, 34, 35), and the presence of interference fringes with high Fisher information, which is useful for quantum-enhanced sensing (4, 36, 37). REFERENCES AND NOTES 1. E. Schrödinger, Naturwissenschaften 23, 807–812 (1935). 2. A. Bassi, K. Lochan, S. Satin, T. P. Singh, H. Ulbricht, Rev. Mod. Phys. 85, 471–527 (2013). 3. L. Pezzè, A. Smerzi, Atom Interferometry, Proceedings of the International School of Physics “Enrico Fermi,” Varenna, G. M. Tino, M. A. Kasevich, eds. (IOS Press, Amsterdam, 2014). 4. W. J. Munro, K. Nemoto, G. J. Milburn, S. L. Braunstein, Phys. Rev. A 66, 023819 (2002). 5. P. T. Cochrane, G. J. Milburn, W. J. Munro, Phys. Rev. A 59, 2631–2634 (1999). 6. M. Mirrahimi et al., New J. Phys. 16, 045014 (2014). 7. C. Monroe, D. M. Meekhof, B. E. King, D. J. Wineland, Science 272, 1131–1136 (1996). 8. H.-Y. Lo et al., Nature 521, 336–339 (2015). 9. A. Ourjoumtsev, H. Jeong, R. Tualle-Brouri, P. Grangier, Nature 448, 784–786 (2007). 10. K. Huang et al., Phys. Rev. Lett. 115, 023602 (2015). 11. A. Auffeves et al., Phys. Rev. Lett. 91, 230405 (2003). 12. S. Deléglise et al., Nature 455, 510–514 (2008). 13. B. Vlastakis et al., Science 342, 607–610 (2013). 14. D. Leibfried et al., Nature 438, 639–642 (2005). 15. W.-B. Gao et al., Nat. Phys. 6, 331–335 (2010). Bild et al., Science 380, 274–278 (2023) 21 April 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E 16. J. R. Friedman, V. Patel, W. Chen, S. K. Tolpygo, J. E. Lukens, 32. B. Schrinski et al., Phys. Rev. Lett. 130, 133604 Nature 406, 43–46 (2000). 17. T. Kovachy et al., Nature 528, 530–533 (2015). 18. Y. Y. Fein et al., Nat. Phys. 15, 1242–1245 (2019). 19. Y. Chu et al., Science 358, 199–202 (2017). 20. A. Blais, A. L. Grimsmo, S. M. Girvin, A. Wallraff, Rev. Mod. Phys. 93, 025005 (2021). 21. See supplementary materials. 22. Y. Chu et al., Nature 563, 666–670 (2018). 23. U. von Lüpke et al., Nat. Phys. 18, 794–799 (2022). 24. V. Bužek, H. Moya-Cessa, P. L. Knight, S. J. D. Phoenix, Phys. Rev. A 45, 8190–8203 (1992). 25. F. W. Cummings, Phys. Rev. 140, A1051–A1056 (1965). 26. G. Rempe, H. Walther, N. Klein, Phys. Rev. Lett. 58, 353–356 (1987). 27. J. H. Eberly, N. B. Narozhny, J. J. Sanchez-Mondragon, Phys. Rev. Lett. 44, 1323–1326 (1980). 28. J. Gea-Banacloche, Phys. Rev. Lett. 65, 3385–3388 (1990). 29. K. S. Chou et al., Nature 561, 368–373 (2018). 30. A. Grimm et al., Nature 584, 205–209 (2020). 31. A. Kenfack, K. Zyczkowski, J. Opt. B Quantum Semiclassical Opt. 6, 396–404 (2004). (2023). 33. C. Whittle et al., Science 372, 1333–1336 (2021). 34. N. Ofek et al., Nature 536, 441–445 (2016). 35. C. Flühmann et al., Nature 566, 513–517 (2019). 36. J. Joo, W. J. Munro, T. P. Spiller, Phys. Rev. Lett. 107, 083601 (2011). 37. A. Facon et al., Nature 535, 262–265 (2016). 38. M. Bild et al., Schrödinger cat states of a 16-microgram mechanical oscillator, Zenodo (2023); https://doi.org/10.5281/ zenodo.7701912. ACKN OWLED GMEN TS We thank O. Romero-Isart, A. Grimm, and I. C. Rodrigues for useful discussions, and A. Brooks for help with figure making. Fabrication of devices was performed at the FIRST cleanroom of ETH Zürich and the BRNC cleanroom of IBM Zürich. Funding: M.B. was supported by the QuantERA II Program that has received funding from the European Union’s Horizon 2020 research and innovation program under grant agreement no 101017733, and with the Swiss National Science Foundation. M.F. was supported by The Branco Weiss Fellowship–Society in Science, administered by the ETH Zürich. P.M. was supported by the ThinkSwiss fellowship. Author contributions: U.v. L. fabricated the device. M.B., M.F., Y.Y., and U.v. L. performed the experiments and analyzed the data. All authors performed theoretical calculations and simulations. Y.C. conceived of the project and supervised the work. M.B., M.F., Y.Y., U.v. L., and Y.C. wrote the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: All data and software are available in the manuscript or the supplementary material or are deposited at Zenodo (38). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf7553 Supplementary Text Figs. S1 to S8 Tables S1 and S2 References (39–46) Submitted 10 November 2022; accepted 20 March 2023 10.1126/science.adf7553 Bild et al., Science 380, 274–278 (2023) 21 April 2023 5 of 5
10.1126_science.adf9728
RES EARCH MESOSCOPIC PHYSICS Universal chiral Luttinger liquid behavior in a graphene fractional quantum Hall point contact Liam A. Cohen1†, Noah L. Samuelson1†, Taige Wang2,3, Takashi Taniguchi4, Kenji Watanabe5, Michael P. Zaletel2,3, Andrea F. Young1* One-dimensional conductors are described by Luttinger liquid theory, which predicts a power-law suppression of the single-electron tunneling density of states at low voltages. The scaling exponent is predicted to be quantized when tunneling into a single isolated chiral edge state of the fractional quantum Hall effect. We report conductance measurements across a point contact linking integer and fractional quantum Hall edge states (at fillings 1 and 1 the predicted universal quadratic scaling with temperature and voltage. At strong coupling, we demonstrate perfect Andreev reflection of fractionalized quasiparticles at the point contact. We use the strong coupling physics to realize a nearly dissipationless direct current voltage step-up transformer, whose gain arises directly from topological fractionalization of electrical charge. 3, respectively). At weak coupling, we observe T he Landau theory of Fermi liquids pro- vides a near-ubiquitous description of interacting fermion systems. One excep- tion is provided when electrons are con- fined to one dimension, where arbitrarily weak interactions favor a distinct phase known as the Tomonaga-Luttinger liquid (1–4). In this phase, the low-energy collective excitations are orthogonal to the single-electron operators from which they are microscopically constructed. This “orthogonality catastrophe” manifests experimentally as a power-law suppression of the electron tunneling density of states, (cid:3) (cid:1) (cid:2)1 N Eð ÞºðE (cid:2) EFÞ , at the Fermi energy EF, even though the system remains conductive. The power law is characterized by an ex- ponent g known as the Luttinger parameter, which depends continuously on the nature and strength of the interparticle interactions (5). Experimentally, Luttinger liquid behavior can manifest through a non-Ohmic current- voltage relation I Vð ÞºV g , as observed, for example, in ropes of single-walled carbon nanotubes (6, 7). 1 g 1 Chiral Luttinger liquids An alternative means of creating a one- dimensional (1D) wire can be found at the boundary of a topologically ordered phase, as occurs in the fractional quantum Hall (FQH) effect (8). Here, the right- and left- moving modes are physically separated to opposite edges of the 2D sample, resulting in a chiral Luttinger liquid in which backscattering is suppressed entirely and the Luttinger param- 1Department of Physics, University of California at Santa Barbara, Santa Barbara, CA 93106, USA. 2Department of Physics, University of California, Berkeley, CA 94720, USA. 3Material Science Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA. 4International Center for Materials Nanoarchitectonics, National Institute for Materials Science, 1-1 Namiki, Tsukuba 305-0044, Japan. 5Research Center for Functional Materials, National Institute for Materials Science, 1-1 Namiki, Tsukuba 305-0044, Japan. *Corresponding author. Email: andrea@physics.ucsb.edu †These authors contributed equally to this work. eter g becomes quantized (9). In this setting, g becomes a fingerprint of the topological order of the enclosed bulk; for example g ¼ 1 3 for the Laughlin state at Landau level filling n ¼ 1 3 (9, 10). In such a scenario, tunneling of whole electrons into the fractionalized edge of the n ¼ 1 3 state is predicted to exhibit a quadratic scaling of the tunneling conductance G ≡ dI dV with both the temperature T and bias voltage V: G º T2, V2. The quadratic scaling is a direct result of the quantized value of g ¼ 1 3, and a central prediction of the theory is that this behavior is universal and independent of mi- croscopic details at sufficiently low tempera- tures and bias voltages (9). Experimentally, one of the most successful tests of the chiral Luttinger liquid theory was obtained by studying tunneling between the edge of a 2D-electron gas (2DEG) hosting a FQH state at n ¼ 1 3 and a 3D electrode grown on the cleaved edge of the semiconductor wafer (3, 11). Although striking power-law behavior was observed over a wide range of bias voltages and temperatures (11), the ex- ∼2:7 was found to vary across sam- ponent 1 g 3 plateau (12, 13)—in ples and within the n ¼ 1 disagreement with the predicted quantized ¼ 3. This sparked a thorough dis- exponent 1 g cussion about the nature of bulk-boundary correspondence and whether g is indeed a distinct imprint of the topological order of the bulk (14–16). However, the possibility of edge reconstruction was soon identified as a con- founding factor that could reconcile the range of observed exponents with the chiral Luttinger liquid theory (17–19). In principle, a quantum point contact be- tween integer and fractional quantum Hall edge states (5, 9, 10, 20, 21–23) provides a richer test bed for chiral Luttinger liquid physics. In this geometry, the collective modes of the 3 and n = 1 edges may be modeled as n ¼ 1 chiral bosonic fields fa and fb, respectively, coupled by a single point scatterer of strength G. The low energy physics are described by the Lagrangian L ¼ 1 4p X @xf i ð @t (cid:2) @x Þf i þ i¼a;b (cid:1) (cid:3) b b a a ½ p y ð1Þ (cid:3) ¼ 1 ¼ ei Gd xð Þ y† a ½ 2 whereas y þ y† y b ffiffi ¼ eifb fa and y where the operators y 3 3 and n = 1 edges, remove an electron on the n ¼ 1 respectively. In Eq. 1, the first term describes the gapless bosonic edge modes on either side of the junction, whereas the second term de- scribes inter-edge tunneling of electrons at the point contact. In the language of the renor- malization group, the scaling dimensions of (cid:3) ¼ 3 the electron operator y b 2 is three times larger, reflecting the topological order of the n ¼ 1 3 fractional quantum Hall bulk. For the corresponding 2D Euclidean ac- (cid:3)(cid:2) tion to remain dimensionless, G½ (cid:3) ¼ (cid:2)1, meaning that edge-to-edge tunnel- ½ y ing is irrelevant and becomes increasingly less important at low energies. This leads to the conclusion that no matter how “open” the junc- tion is made—in other words, no matter how large the bare value of G is— the conductance will vanish at a sufficiently low temperature and voltage bias (24). Near this decoupled fixed point, the conductance can be computed with- in perturbation theory (3), giving (cid:3) ¼ 1 (cid:2) y ½ a a b Þ (cid:5) ð G V ; T e2 2h ¼ 2pT T0 (cid:6) " 2 1 3 þ (cid:5) eV 2pkbT (cid:6) 2 # þ … ð2Þ for small temperatures T and voltage bias V. Here, the bare tunneling strength is repre- sented by T0, where for weakly coupled edges T0º 1 G. In this limit, the power law exponent (cid:8) (cid:3) (cid:7)(cid:1) ¼ 2 describing the T and V depen- (cid:2) 1 1 g dence provides a direct probe of the bulk topo- logical order. Quantum point contact experiments in semi- conductor quantum wells have indicated power- law behavior near the decoupled fixed point (close to full pinch-off) over a limited temper- ature range (25, 26); however, the presence of disorder at the tunnel junction often compli- cates the physical interpretation (25–31). In a more recent experiment (32) focused on the weak quasiparticle backscattering limit, the conductance showed clear characteristics of a Luttinger liquid but remained quantitatively inconsistent with the predictions of chiral Luttinger liquid theory: The measured value of the Luttinger parameter g did not align with predictions for any candidate incompressible ground states at the filling factors studied in (31). It appears that for both quantum point contact and cleaved-edge overgrowth experi- ments, nonuniversal effects arising from the de- tailed structure of experimentally realized edges Cohen et al., Science 382, 542–547 (2023) 3 November 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Universal conductance scaling at weak coupling. (A) Optical micrograph of the device with schematic depiction of chiral edge states. (B) Device schematic showing the patterned top graphite layer, graphene monolayer, and global bottom graphite gate. The tunneling conductance across the junction is determined from the transmitted current I and diagonal voltage VD as G (cid:4) I VD (C) G measured as function of V at Tprobe = 56 mK, with VNS = −2.465 V and B = 10 T. (Inset) parabolic fit to the low-V regime, giving T0 = 9.02 ± 0.007 K as defined in Eq. 2. The main panel shows G − Gmin, where . (cid:2)4 e2 h is the minimum conductance. Fitting a power law gives an Gmin ¼ 7:5 (cid:5) 10 exponent of 2.00 ± 0.06 (44), where the error represents the standard deviation in the fit parameter. (D) G measured at V = 0 as a function of temperature at the same gate voltages as (C). The dashed line is a plot of the conductance given by the first term of Eq. 2, using T0 = 9.02 K. (E) Nonlinear differential conductance for Tprobe = 202, 245, 290, 344, 450, 549, 618, and 666 mK at VNS = −2.456 V. (F) The same data as in (E) after scaling G and V as described in the text. The black dashed line is G˜ as predicted by Eq. 2. (22, 33, 34, 35) push the universal scaling regime to energy scales beyond experimental reach. Conductance scaling in graphene quantum point contact heterojunction Recently, graphene has emerged as a useful platform for precision tests of chiral Luttinger liquid physics. Notably, the low disorder avail- able in field effect–controlled devices make a wide range of fractional quantum Hall states accessible (36, 37). However, prior studies of quantum point contacts have been limited to regimes dominated by resonant tunneling effects, owing at least in part to disorder introduced during the fabrication of local split gates (38–41). We use anodic oxidation lithography to pattern nanoscale features in graphite gates, which are then incorporated into a van der Waals hetero- structure to produce a clean quantum point contact (42). A micrograph and schematic of our device is shown in Fig. 1, A and B. 44-nm- thick hexagonal boron nitride (hBN) dielectric layers are used as spacers between the mono- layer graphene layer and the top and bottom A ) h / 2 e ( G B V r V i C i V / r V 1.0 0.5 0.0 -0.5 0.5 1/2 1/3 0.0 -3.4 VNS (V) -3.0 -e* 1/3 1 2e* e 0 -1/2 -3.4 VNS (V) -3.0 3 on the east and west Fig. 2. Andreev-like quasiparticle scattering. (A) G versus VNS taken at B = 9 T with a finite DC voltage bias of 145 µV. Here, VBG = 2.0 V, VE = −1.460 V, and VW = −1.775 V to maintain n = 1 and n ¼ 1 sides of the junction, respectively. (B) Schematic representation of the Andreev scattering process for fraction- ally charged quasiparticles in the strong coupling limit (23) of a n ¼ 1 3 to n = 1 point heterojunction. (C) Ratio of the reflected voltage Vr to the source voltage Vi versus VNS; all other gate voltages are the same as in (A). Cohen et al., Science 382, 542–547 (2023) 3 November 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Crossover from weak to strong coupling. (A) Schematic of renormalization group flow in the G−U plane, where U represents additional perturbations to Eq. 1. (B) G as a function of the voltage on the north and south gates, VNS, and the DC voltage bias V at Tprobe = 56 mK and B = 10 T. (C) Line cuts of the data in (B) at the values of VNS indicated by the colored points. Black dashed lines are plotted using the value of G predicted by Eq. 3 where the parameter T0 is extracted from the low-bias conductance. (D) The zero-bias conductance also scales with temperature in agreement with Eq. 3 for low energies. Although the data deviate from the model at high energies, G nevertheless exceeds G ¼ e2 3h for T >> T0, indicating strong coupling at high T. (E) The data from (D), after scaling T by T0. The curves collapse onto the universal scaling formula Eq. 3, shown in black. gates. The device architecture features a four- quadrant split gate geometry (42) where independent voltages may be applied to the north, south, east, and west top gates (VN, VS, VE, and VW, respectively) as well as a global bottom gate (VBG). With VBG held constant, VE and VW fix the filling factors of the east and west regions to 1 and 1 3, respectively, whereas VNS = VN = VS is used to create a constriction by tuning the filling of the north and south re- gions to n ≤ 0. The choice of VBG controls the “sharpness” of the potential profile at the constriction, pa- rameterized by the energy scale EV ≡ e @V ‘b. @x Here V, which is defined as the applied po- tential, is purely a function of the applied gate voltages and does not take into account the finite compressibility of the 2DEG at a high magnetic field. EV plays a key role in the physics of fractional quantum Hall edges: When the confinement energy EV is smaller than the Coulomb energy, the edge may re- construct (33, 43), introducing additional edge modes which may push the universal tunneling behavior to experimentally inaccessible energy scales. As described in (42) and (44), the control available in our geometry allows us to access the universal regime of EC < EV while main- taining independent control of the bulk filling factor and quantum point contact transparency. We begin by investigating the “weak coupl- ing” regime where G ≪ e2 h . This corresponds to the limit of eV ≪ 2pkbT0 and T ≪ T0 2p where Eq. 2 holds. We measure G ≡ I (see Fig. 1B), which is VD a four-terminal measurement of the tunneling conductance (42, 44). We tune T0 through VNS. Figure 1C shows G measured as a func- tion of the direct current (DC) voltage bias V at a fixed Tprobe = 56 mK and VNS = −2.465 V. As seen in the inset, the V dependence is well fit by a parabola, with a curvature corresponding to T0 = 9.02 K in Eq. 2. To assess the quality of the power law fit, we subtract the minimum conductance Gmin and plot G − Gmin on a logarithmic scale (Fig. 1C). We find a simple V2 power law over one order of magnitude in V, corresponding to two orders of magnitude in G − Gmin. Specifically, we find an exponent of 2.00 ± 0.06. Although we have not studied the filling factor dependence in detail, fig. S7 shows a second measurement taken at a different magnetic field and different value of the bulk filling factor within the n ¼ 1 3 plateau, which yields the same exponent (44). We next compare experiment with the pre- dicted zero-bias temperature dependence, G (V = 0, T, T0) of Eq. 2. The result for T0 = 9.02 K is shown in Fig. 1D. In contrast to the volt- age dependence, a T2 power law is observed only for a limited range of T, between 200 and 700 mK. We attribute deviations at lower tem- peratures to a decoupling of the electronic tem- perature from Tprobe. At high temperatures, deviations are expected as corrections to Eq. 2 become important, with significant deviations 4p ≈ 750mK (44). Although onsetting for T > T0 the power law behavior in T occurs over a lim- ited range, the observed power law in T is con- sistent with the more robust power law in V, Cohen et al., Science 382, 542–547 (2023) 3 November 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E which is a manifestation of a general scal- (cid:1) (cid:3)2⋅ h ing relation. Defining ~G ¼ 2G⋅ T0 e2 and (cid:1) (cid:3) 2pT Þ, it follows that ~G ¼ 1 x ¼ eV þ x2 þ ⋯ ð 2pkBT provides a universal low-energy collapse. 3 Figure 1E shows G as a function of V for sev- eral different temperatures at VNS = −2.456 V. The same data, plotted as G̃ (x), are shown in Fig. 1F. In these datasets, T0 = 6.87 K is de- termined by fitting the lowest temperature curve in Fig. 1E to Eq. 2. For low values of V where universality is expected (44), the curves collapse onto the universal parabolic curve expected from chiral Luttinger liquid theory. From weak to strong coupling As seen in Fig. 1E, G approaches e2 2h at high bias. One might expect that full transmission of the incoming fractional edge mode would cause the conductance to saturate at G ¼ e2 3h. In fact, the observed G≈ e2 2h can be understood from the peculiar properties of the point contact at strong coupling, defined as eV ≫kbT0 or T ≫T0. The strong coupling regime is accessed most readily by lowering T0 (i.e., increasing VNS) at fixed V and T, leading to a plateau in the dif- ferential G≈ e2 2h (Fig. 2A). Microscopically, the excess conductance can be understood by analogy to Andreev scattering at a metal-superconductor interface (22, 23, 45). In this picture, conduction occurs when a pair of incident quasiparticles, each with charge e(cid:6) ¼ (cid:2) e 3, is transmuted into a single electron on the n = 1 side through the simultaneous retro- reflection of a charge e 3 quasi-hole into the downstream chiral edge state (see Fig. 2B). This process leads to the observed nearly quantized increase in G. Moreover, when the Andreev pro- cess is the dominant form of charge transfer across the junction, the downstreamn ¼ 1 3 edge hosting the retroreflected hole is expected to develop a negative chemical potential with mag- nitude one-half of the voltage of the incoming edge (23, 46). Figure 2C shows the measured reflected voltage Vr under the same conditions as the G data in Fig. 2A. The near-quantization of both G and Vr over the same broad range of Vi junction transparency implies that the described Andreev process completely dominates charge transfer. This contrasts with previous experi- ments in semiconductor wells (30), where a much smaller effect was observed, likely me- diated by resonant scattering. In our discussion, we have considered only single-electron tunneling between edges, pa- rameterized by G; however, additional pro- cesses such as electron co-tunneling may also contribute, which would be represented by additional operators not included in Eq. 1. Re- normalization group analyses have shown (10) that as G → 0 these terms are more irrelevant than G. This guarantees that regardless of mi- croscopic details Eq. 1 becomes a good approx- imation to the physical system at sufficiently low energies. This is characteristic of a stable Fig. 4. Zero frequency voltage step-up trans- former. (A) The differen- tial gain dVo and resulting dVi integrated DC gain b ¼ dVo , dVi measured in the config- uration shown in the inset, with B = 9 T, Tprobe = 48 mK, VE = −1.460 V, VW = −1.775 V, VNS = −3.225 V, VBG = 2.0 V. The FQH Andreev scattering process yields an enhancement of the output voltage on the FQH side (46), with the DC gain predicted to reach a value of 1.5 in the dis- sipationless limit. Experi- mentally, we find a gain b ¼ 1:46, despite the nonlinearity at low bias arising from the sup- pression of the Andreev scattering at low energies. (B) The DC power dissipation ratio, calculated from b through Pout=Pin ¼ ð2b þ 1, is plotted versus V, and reaches a maximum value of 97.6% (49). (cid:9) (cid:10) Þ2 (cid:2) 2b 3 3 fixed point of the renormalization group and accounts for the universal scaling behavior demonstrated in Fig. 1. A different result is obtained at strong cou- pling, formally G → ∞, which represents an unstable fixed point (23, 47). At this fixed point the dominant process that transfers charge across the junction is Andreev scattering. How- ever, while other processes vanish when G = ∞, giving G ¼ e2 2h, for any finite value of G, G is expected to vanish at low energies as the sys- tem flows toward the stable G = 0 fixed point. This is shown schematically in Fig. 3A, which depicts a renormalization group flow diagram indicating the trajectory of the conductance as the energy is lowered. In this plot, the y-axis, G, represents the coefficient of the operator which transfers an electron between the two edge modes, whereas the x-axis represents the coefficient “U” of any operator not explicitly captured in Eq. 1. As is seen in the diagram, finite U, expected for realistic heterojunctions, would seem incom- patible with approaching the strong coupling fixed point. For this reason, the strong coupling limit was previously thought to be practically in- accessible except through highly tuned resonant scattering (46, 48). Although fine-tuned, this limit would arise if the QPC forms an adiabatic constriction in which approximate momentum conservation along the QPC prevents backscat- tering between the N and S edge (49). The fact that we observe near-perfect Andreev reflection when the junction is highly transmis- sive suggests that the microscopics of the system approach the strong coupling fixed point with negligible contributions from the additional operators represented by U. When U = 0, Eq. 1 can be mapped to an integrable “quantum impurity model” with an exact solution for all values of G (24, 47). The solution provides an expression for G at arbitrary V, T, and T0, ð G V ; T Þ ¼ ∫∞ (cid:2)∞ e2 2h ð Þ2 (cid:2) 2E þ eV Þ2 þ kbT0 ð ð 2E þ eV Þ2 f ′ Eð ÞdE ð3Þ where f(E) is the Fermi-Dirac function (47). Equation 3 provides a universal crossover function which describes the transition be- tween weak and strong coupling, depicted as a single line along the y-axis (U = 0) of Fig. 3A (22, 23, 47). The integral reduces to Eq. 2 for T0 → ∞, corresponding to the weak coupling fixed point, and gives e2 2h for T0 = 0 corresponding to strong coupling. Figure 3B shows G as a function of V and VNS at a probe temperature Tprobe = 56 mK. Throughout the plotted range, the high-V conductance saturates to approximately e2 2h as expected for the strong coupling fixed point. Meanwhile, a zero-bias dip remains visible for all VNS, indicative of the instability of the point contact to edge decoupling at low ener- gies. To quantify how closely the quantum im- purity model describes our system, we compare Eq. 3 with our experimental data as a function of V and T, and T0. In Fig. 3C, we plot four curves extracted for different values of VNS. For each curve, T0 is determined from a fit to the low bias behavior with an appropriate low- energy expansion of Eq. 3 (44). For the largest fit values of T0, the residual value of G when V = 0 Cohen et al., Science 382, 542–547 (2023) 3 November 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E can then be used as a primary thermometer—in effect allowing us to correct for a possible lack of equilibration between the electron temper- ature and the probe thermometer. Using this method, we find Telectron = 91 mK for the T0 = 9.02 K trace, in contrast to the measured probe temperature of 56 mK. Taking 91 mK as the electron temperature for the remaining data sets in Fig. 3C, Eq. 3 may then be used to generate V- dependent curves interpolating between weak and strong coupling. These curves are overlaid in black on the experimental data in Fig. 3C. Figure 3D shows G (V = 0, T) plotted as a function of the probe temperature for the same values of VNS as in Fig. 3C, along with the results of Eq. 3 for the corresponding values of T0. Although Eq. 3 does not provide a simple scaling between temperature and voltage as is available at the weak coupling fixed point, for V = 0 the conductance can be written as a function of the scaled temperature, T . Figure T0 3E shows the V = 0 conductance plotted as a function of T . The four data sets shown in T0 unscaled form in Fig. 3E collapse onto different parts of the universal crossover function be- tween weak and strong coupling. Note that for Fig. 3, D and E, the temperature is measured on the probe and no corrections are made for disequilibrium between Telectron and Tprobe, leading to systematic deviations between experiment and theory at the lowest tempera- tures. The collapse of the experimental data onto a single universal curve over nearly two orders of magnitude in T0 strongly supports the conclusion that our quantum point contact realizes the exactly solvable Lagrangian of Eq. 1 to a high degree of accuracy, and in particular, that tunneling predominantly occurs through a single point at the QPC rather than by multiple disorder-induced scattering centers. Near-dissipationless step-up transformer At the strong coupling fixed point the quantum point contact acts as a nearly dissipationless splitter, partitioning current injected on the n = 1 edge equally between the downstream n ¼ 1 3 and the upstream n = 1 edge states (23). For G = ∞, the partitioning happens with unit proba- bility. In this limit, no entropy is generated and dissipation does not occur, leading to unity power efficiency (46). This contrasts with the more conventional case of partial transmission of edge modes at a quantum point contact, where fluctuations arising from partition noise lead to dissipation. The dissipationless current splitting property of the strong coupling fixed point allows us to operate our device as a voltage step-up transformer (23, 46). Figure 4 shows the differential gain, dVo , where Vi is the dVi input voltage applied to the upstream n = 1 edge state and Vo is the output voltage measured on the downstream n ¼ 1 3 edge state with all other contacts grounded. The differential gain is very close to the theoretically expected value of 3 2 (46). Because it is built on a purely zero fre- quency effect, the transformer gain remains al- most the same even in the “direct current” (DC) limit, as can be seen in the behavior of the integrated gain, Vo , in comparison to its differ- Vi ential counterpart. The exceptionally high power efficiency, which peaks at 97.6%, is a testament to how closely the experiment realizes the strong-coupling fixed point and contrasts fa- vorably with zero frequency voltage amplifi- cation based on superconductors (50, 51) and bilayer quantum Hall systems (49, 51). Discussion and outlook Our observation of universal chiral Luttinger liquid physics at both weak and strong coupl- ing directly paves the way for experiments on two dimensional systems where mesoscopic electrostatic control plays a key role in address- ing unanswered questions about strong cor- relations, topological order, and quantum statistics. Examples include even denominator fractional quantum Hall states observed in mono- (52, 53) and bilayer (38, 54, 55) gra- phene, where taming edge reconstruction in a quantum point contact may allow for un- ambiguous experimental constraints on the ground state topological order (45, 56–59). In addition, the flexibility inherent in anodic oxidation prepatterning will allow independent tuning of quasiparticle number, edge sharp- ness, and quantum point contact transparency in interferometer geometries where direct access to quantum statistics is possible (60, 61). Finally, gate-defined point contacts may allow for precision measurements of order parame- ters in the recently discovered crystalline graphene superconductors (62–64). RE FERENCES AND NOTES J. M. Luttinger, J. Math. Phys. 4, 1154–1162 (1963). 1. 2. S.-I. Tomonaga, Prog. Theor. Phys. 5, 544 (1950). 3. A. M. Chang, Rev. Mod. Phys. 75, 1449–1505 (2003). 4. T. Giamarchi, T. Nattermann, P. L. Doussal, Quantum Physics in One Dimension (Clarendon Press, 2003). 5. M. P. A. Fisher, L. I. Glazman, Mesoscopic Electron Transport, L. L. Sohn, L. P. Kouwenhoven, G. Schön, eds., NATO ASI Series (Springer, 1997), pp. 331–373. 6. M. Bockrath et al., Nature 397, 598–601 (1999). 7. H. Ishii et al., Nature 426, 540–544 (2003). 8. B. I. Halperin, Phys. Rev. B Condens. Matter 25, 2185–2190 (1982). 9. X. G. Wen, Phys. Rev. B Condens. Matter 41, 12838–12844 (1990). 10. C. L. Kane, M. P. A. Fisher, Perspectives in Quantum Hall Effects (Wiley, 1996), pp. 109–159. 11. A. M. Chang, L. N. Pfeiffer, K. W. West, Phys. Rev. Lett. 77, 2538–2541 (1996). 12. M. Grayson, D. C. Tsui, L. N. Pfeiffer, K. W. West, A. M. Chang, Phys. Rev. Lett. 80, 1062–1065 (1998). 13. A. M. Chang, M. K. Wu, C. C. Chi, L. N. Pfeiffer, K. W. West, Phys. Rev. Lett. 86, 143–146 (2001). 14. V. J. Goldman, E. V. Tsiper, Phys. Rev. Lett. 86, 5841 (2001). 15. S. S. Mandal, J. K. Jain, Phys. Rev. Lett. 89, 096801 (2002). 16. S. Jolad, J. K. Jain, Phys. Rev. Lett. 102, 116801 (2009). 17. X. Wan, F. Evers, E. H. Rezayi, Phys. Rev. Lett. 94, 166804 (2005). 18. Z.-X. Hu, R. N. Bhatt, X. Wan, K. Yang, Phys. Rev. Lett. 107, 236806 (2011). 19. D. Varjas, M. P. Zaletel, J. E. Moore, Phys. Rev. B Condens. Matter Mater. Phys. 88, 155314 (2013). 20. D.-H. Lee, X.-G. Wen, Phys. Rev. Lett. 66, 1765–1768 (1991). 21. I. Safi, Ann. Phys. 22, 463–679 (1997). 22. C. C. Chamon, E. Fradkin, Phys. Rev. B Condens. Matter 56, 2012 (1997). 23. N. P. Sandler, C. C. Chamon, E. Fradkin, Phys. Rev. B Condens. Matter Mater. Phys. 57, 12324–12332 (1998). 24. C. L. Kane, M. P. A. Fisher, Phys. Rev. B Condens. Matter 46, 15233–15262 (1992). 25. F. P. Milliken, C. P. Umbach, R. A. Webb, Solid State Commun. 97, 309–313 (1996). 26. I. Maasilta, V. Goldman, Phys. Rev. B Condens. Matter 55, 4081–4084 (1997). 27. S. Roddaro et al., Phys. Rev. Lett. 90, 046805 (2003). 28. I. P. Radu et al., Science 320, 899–902 (2008). 29. J. B. Miller et al., Nat. Phys. 3, 561–565 (2007). 30. M. Hashisaka et al., Nat. Commun. 12, 2794 (2021). 31. S. Baer et al., Phys. Rev. B Condens. Matter Mater. Phys. 89, 085424 (2014). 32. S. Baer et al., Phys. Rev. B Condens. Matter Mater. Phys. 90, 075403 (2014). 33. Cd. Chamon, X. G. Wen, Phys. Rev. B Condens. Matter 49, 8227–8241 (1994). 34. X. Wan, E. H. Rezayi, K. Yang, Phys. Rev. B Condens. Matter Mater. Phys. 68, 125307 (2003). 35. K. Yang, Phys. Rev. Lett. 91, 25 (2003). 36. A. A. Zibrov et al., Nature 549, 360–364 (2017). 37. C. Dean, P. Kim, J. I. A. Li, A. Young, Fractional Quantum Hall Effects: New Developments (World Scientific, 2020), pp. 317–375. 38. K. Zimmermann et al., Nat. Commun. 8, 14983 (2017). 39. H. Overweg et al., Nano Lett. 18, 553–559 (2018). 40. C. Déprez et al., Nat. Nanotechnol. 16, 555–562 (2021). 41. Y. Ronen et al., Nat. Nanotechnol. 16, 563–569 (2021). 42. L. A. Cohen et al., Tunable fractional quantum Hall point contacts in graphene via local anodic oxidation of graphite gates. arXiv:0706.1234 [cond-mat.mes-hall] (2022). 43. X. Wan, K. Yang, E. H. Rezayi, Phys. Rev. Lett. 88, 056802 (2002). 44. Materials and methods can be found in the supplementary materials. 45. M. Milovanović, N. Read, Phys. Rev. B Condens. Matter 53, 13559–13582 (1996). 46. D. B. Chklovskii, B. I. Halperin, Phys. Rev. B Condens. Matter Mater. Phys. 57, 3781–3784 (1998). 47. P. Fendley, H. Saleur, N. P. Warner, Nucl. Phys. B 430, 577–596 (1994). 48. C. L. Kane, Resonant Tunneling between quantum Hall states at filling \nu = 1 and \nu = 1/3. arXiv:9809020 [cond-mat.mes-hall] (1998). 49. B. I. Halperin, A. Stern, S. M. Girvin, Phys. Rev. B Condens. Matter Mater. Phys. 67, 235313 (2003). 50. I. Giaever, IEEE Spectr. 3, 117–122 (1966). 51. S. M. Girvin, D. C. Transformer, D. C. Josephson, Phys. Scr. 2002, 112 (2002). 52. A. A. Zibrov et al., Nat. Phys. 14, 930–935 (2018). 53. Y. Kim et al., Nat. Phys. 15, 154–158 (2019). 54. D.-K. Ki, V. I. Fal’ko, D. A. Abanin, A. F. Morpurgo, Nano Lett. 14, 2135–2139 (2014). 55. J. Li et al., Science 358, 648–652 (2017). 56. X.-G. Wen, Phys. Rev. Lett. 70, 355–358 (1993). 57. P. Fendley, M. P. A. Fisher, C. Nayak, Phys. Rev. Lett. 97, 036801 (2006). 58. R. Ohashi, R. Nakai, T. Yokoyama, Y. Tanaka, K. Nomura, J. Phys. Soc. Jpn. 91, 123703 (2022). 59. W. Bishara, C. Nayak, Phys. Rev. B Condens. Matter Mater. Phys. 77, 165302 (2008). 60. C. Nayak, S. H. Simon, A. Stern, M. Freedman, S. Das Sarma, Rev. Mod. Phys. 80, 1083–1159 (2008). 61. J. Nakamura et al., Nat. Phys. 15, 563–569 (2019). 62. H. Zhou, T. Xie, T. Taniguchi, K. Watanabe, A. F. Young, Nature 598, 434–438 (2021). 63. H. Zhou et al., Science 375, 774–778 (2022). 64. Y. Zhang et al., Nature 613, 268–273 (2023). 65. L. A. Cohen et al., Universal chiral Luttinger liquid behavior in a graphene fractional quantum Hall point contact, Version 1.0, Zenodo (2022); https://doi.org/10.5281/zenodo.7369945. AC KNOWLED GME NTS The authors acknowledge discussions with B. L. Halperin, C. Nayak, and C. Kane, and A. Assouline and A. M. Potts for helpful comments on the manuscript. Funding: Work at UCSB was primarily supported by the Air Force Office of Scientific Research under award FA9560-20-1-0208 and by the Gordon and Betty Moore Foundation EPIQS program under award GBMF9471. L.C. and N.S. received additional support from the Army Research Office under award W911NF20-1-0082. T.W. and M.Z. were supported by the Director, Office of Science, Office of Basic Energy Sciences, Cohen et al., Science 382, 542–547 (2023) 3 November 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E Materials Sciences and Engineering Division of the US Department of Energy under contract no. DE-AC02-05-CH11231 (van der Waals heterostructures program, KCWF16). M.Z. received additional support from the Army Research Office through the MURI program (grant number W911NF-17-1-0323) K.W. and T.T. acknowledge support from JSPS KAKENHI (Grant Numbers 19H05790, 20H00354 and 21H05233). Author contributions: L.A.C. and N.L.S. fabricated the device and performed the measurements. L.A.C., N.L.S., M.P.Z., T.W., and A.F.Y. analyzed the data and wrote the paper. T.T. and K.W. synthesized the hBN crystals. Competing interests: A.F.Y. is a board member of sp2 quantum. L.A.C. and A.F.Y. are inventors on the patent application PCT/ US23/14059, which is related to this work. Data and materials availability: All of the experimental data presented and code for generating the figures from the datasets are available in a public Zenodo repository (65). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/ science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf9728 Materials and Methods Supplementary Text Figs. S1 to S10 Reference (66) Submitted 29 November 2022; resubmitted 17 December 2022 Accepted 29 September 2023 10.1126/science.adf9728 Cohen et al., Science 382, 542–547 (2023) 3 November 2023 6 of 6
10.1126_science.adf5568
RES EARCH CHROMATIN BIOPHYSICS Stochastic motion and transcriptional dynamics of pairs of distal DNA loci on a compacted chromosome David B. Brückner1†, Hongtao Chen2,3†, Lev Barinov3,4, Benjamin Zoller5,6, Thomas Gregor3,5,6* Chromosomes in the eukaryotic nucleus are highly compacted. However, for many functional processes, including transcription initiation, the pairwise motion of distal chromosomal elements such as enhancers and promoters is essential and necessitates dynamic fluidity. Here, we used a live-imaging assay to simultaneously measure the positions of pairs of enhancers and promoters and their transcriptional output while systematically varying the genomic separation between these two DNA loci. Our analysis reveals the coexistence of a compact globular organization and fast subdiffusive dynamics. These combined features cause an anomalous scaling of polymer relaxation times with genomic separation leading to long-ranged correlations. Thus, encounter times of DNA loci are much less dependent on genomic distance than predicted by existing polymer models, with potential consequences for eukaryotic gene expression. L iving systems are built based on informa- tion encoded in chromosomes confined in each cell’s nucleus. These meter-long DNA polymers must be highly compacted to fit into the micrometer-sized structure (1, 2). At the same time, for cells to function, chromosome organization must allow the in- formation content to be accessed and read out through transcription (3, 4). Often, transcrip- tion can only occur through the spatial inter- action of DNA loci such as enhancers and promoters, which find each other dynamically and remain in physical proximity (5–8). Al- though the distances over which many en- hancers function in higher eukaryotes can be up to mega–base pairs in genomic separa- tion (9–12), it is unknown how these elements come into proximity, what their typical dis- tance is in three-dimensional (3D) space, and how they explore this space dynamically in the process. Specifically, it remains unclear how the real-time physical motion of such coupled pairs of DNA loci determines transcriptional encounters and how this depends on their genomic separation. Over the past decade, the advent of chromo- some capture and imaging methods (13) has given key insights into the 3D spatial orga- nization of chromosomes, with the discovery of structural features such as topologically associating domains (TADs) (14–17), phase- separated nuclear condensates (18–20), and larger-scale compartments (21, 22). These or- 1Institute of Science and Technology, Am Campus 1, Klosterneuburg, Austria. 2School of Life Science and Technology, ShanghaiTech University, Shanghai, China. 3Lewis-Sigler Institute for Integrative Genomics, Princeton University, Princeton, NJ, USA. 4Memorial Sloan Kettering Cancer Center, New York, NY, USA. 5Joseph Henry Laboratories of Physics, Princeton University, Princeton, NJ, USA. 6Department of Developmental and Stem Cell Biology, CNRS UMR3738 Paris Cité, Institut Pasteur, Paris, France. *Corresponding author. Email: tg2@princeton.edu †These authors contributed equally to this work. ganizing structures have key implications for transcriptional regulation (23), but they are not static. Rather, they have been revealed to be heterogeneous across cells (24, 25) and dy- namic and short lived in time (26, 27). The role of the real-time dynamics of pairs of loci is only beginning to be understood and remains elusive for focal contacts that are key to es- tablishing enhancer–promoter interactions in many systems (28). Similarly, from a polymer physics perspec- tive, there is a gap in our understanding of the static and dynamic properties of chromosomes. At large scales, across tens to hundreds of TADs, chromosome organization has been shown to be highly compacted in a space-filling configuration (22, 29, 30). A useful null model for this configuration is the crumpled chain (also referred to as fractal globule) with fractal dimension three (22, 31–33). However, the real-time dynamics of DNA loci revealed by live-imaging experiments exhibit subdiffu- sion with exponents in the range of 0.5 to 0.6 (26, 27, 34–36), close to the predictions of the simple Rouse polymer model, which predicts a loosely packed ideal chain polymer configura- tion with fractal dimension two that is in contrast to the compacted architecture of the crumpled chain model. A promising technique to address this gap are scaling approaches that combine fractal organization and subdiffusive dynamics (37–39), but these have never been tested experimentally. Thus far, experimental datasets have given insight into static organization (14–17, 22, 30, 40), dynamic properties of chromosomes (26, 27, 34, 35, 41), or transcription (8, 36, 42–44), but rarely all at the same time. For instance, pre- vious live measurements of locus pairs occurred at fixed genomic separation in transcription- ally silent loci (26, 27). To investigate how 3D spatial organization and dynamic locus mo- tion control the encounter times of functional DNA loci and thus transcriptional activation, we require an approach to simultaneously monitor the movement of DNA locus pairs and transcription across a series of genomic separations in vivo. Here, we addressed this problem by live im- aging the joint dynamics of two cis-regulatory DNA elements, an enhancer and a promoter, while monitoring the transcriptional output resulting from their functional dynamic en- counters in developing fly embryos. We sys- tematically varied the genomic separation between these loci spanning many TADs. Stochastic real-time trajectories of the 3D motion of the two loci showed a dynamic search process, with physical proximity re- quired for successful transcription and a power-law scaling of transcription probability with genomic separation. Although typical 3D distances between the locus pair follow a com- pact packing consistent with the crumpled chain model, the dynamic properties exhibit fast diffusion, albeit with a diffusion coeffi- cient that increases with genomic separation. These features give rise to an anomalous scaling of polymer relaxation times and long- range correlations in the relative motion of the two loci. This suggests that the enhancer- promoter search process is much less depen- dent on genomic separation than expected based on existing polymer models. Results Live imaging of chromosome dynamics and transcription To simultaneously monitor the coupled mo- tion of enhancer-promoter pairs and trans- cription across multiple genomic separations, we generated fly lines in which a reporter gene was introduced at various genomic locations from the well-studied Drosophila even-skipped (eve) locus (8). The locations of both the endo- genous eve enhancers and the promoter of the reporter gene, as well as the transcriptional activity of the reporter gene, were measured together using a three-color imaging system (see the materials and methods, section 1.2, and Fig. 1A) (8). To facilitate transcription, the reporter cassette contained the insulator ele- ment homie, which allowed stable loop forma- tion with the endogenous homie element in the eve locus (Fig. 1B). We built seven such reporter constructs, with genomic separations s varying over close to two orders of magnitude from 58 kb to 3.3 Mb, comparable to the distances over which many enhancers function in higher eukaryotes (see the materials and methods, section 1.1) (9–12). These genomic length scales span across multiple TADs in the Dro- sophila genome, with typical median TAD sizes of ~90 kb (45) (here, 18 kb for the eve locus). Imaging took place for ~30 min during the second half of nuclear cycle 14 (NC14) of Brückner et al., Science 380, 1357–1362 (2023) 30 June 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Simultaneous tracking of DNA loci and transcriptional activity in living embryos. (A) Typical surface view of a representative fly embryo displaying fluorescent foci for MS2, parS, and PP7 in the corresponding blue (top), green (center), and red (bottom) channels. Top inset shows schematic with image location in the embryo; bottom inset shows a close-up. (B) Top: schematic of the gene cassettes used for three-color imaging. The endogenous eve locus (left) is tagged with MS2 stem loops that are labeled with blue fluorescence. A reporter with an eve promoter driving PP7 transcription (labeled with red fluorescence) is integrated at a genomic separation s from the eve locus on the second chromosome in the Drosophila genome. It includes a homie insulator sequence allowing loop formation through homie-homie pairing and a parS sequence that is permanently labeled with green fluorescence. Seven such constructs were generated with varying genomic separation s (triangles). Bottom: sample interlocus distance trajectories R(t) for six genomic separations, with standardized y-axis limits (0, 2 mm) and x-axis limits (0, 30 min) obtained after nucleus and locus segmentation, tracking, chromatic aberration, and motion correction (see the materials and methods, section 1). The sampling time interval is 28 s. (C) Trajectories of interlocus distance R and transcriptional activity, with inferred topological states shown by the colored top bar (blue, Ooff; cyan, Poff; red, Pon; see the materials and methods, section 2). Inset: schematic of the three topological states. (D) 200 examples of state trajectories sampled from a total set of N = 579 trajectories acquired in n = 30 embryos (genomic separation s = 149 kb). Colors are as in (C). Gray trajectory parts correspond to untrackable time points. embryo development (Fig. 1C), well after the completion of DNA replication. Sister chro- matids are tightly coupled together at inter- vals <10 kb (46). Therefore, our two tagged DNA loci are connected by a single chromatin polymer composed of two coupled chromatids that were not resolved by our microscopy. Accordingly, our measurements are associ- ated with increased localization uncertainty and reflect both intra- and interchromosomal interactions, which may not be fully represen- tative of pure intrachromosomal interactions. Interlocus distance scaling suggests a space-filling organization In previous work using a single fixed genomic distance (s = 149 kb) (8), this system was shown to exhibit three topological states (Fig. 1C): an open configuration, Ooff, in which the homie elements are not bound to each other, and two paired configurations, Poff and Pon, in which a loop is formed with either inac- tive or active transcription, respectively. As- suming that these configurations apply to all genomic distances, we determined the in- stantaneous topological and transcriptional states of the system. To this end, we used an inference approach with a hidden Markov model that is based on the time series of inter- locus distances and transcriptional activity (see the materials and methods, section 2). We assigned one of these states to each mea- sured configuration, including the hidden Poff state (Fig. 1D). A key question is how the interlocus dis- tances R in the open configuration Ooff vary with the linear genomic separation s. These distances exhibit broad distributions, which shift systematically with larger separation (Fig. 2A). From a polymer physics perspec- tive, the mean distance hRi is expected to scale as s1/d, where d is the fractal dimension. Whereas an ideal chain polymer, as predicted by the simple Rouse model, has fractal dimen- sion d = 2, the compact crumpled chain or- ganization has dimension d = 3 (33, 47). Our experiments show a scaling exponent of 1/d = 0.31 ± 0.07 for genomic separations up to s = 190 kb, consistent with the crumpled chain model (Fig. 2B). The smaller-than-expected average distances observed for the largest sep- arations (s = 595 kb, 3.3 Mb) are most likely affected by the average folding of the chromo- some (48). The distances of the paired configurations were independent of genomic separation, as anticipated, and exhibited typical distances of 350 to 400 nm (Fig. 2B), consistent with pre- vious measurements of distances within the eve locus (8, 49). Together, these results re- veal a compact crumpled chain architecture of chromosome configurations in a range of genomic separations consistent with Hi-C ex- periments in Drosophila (17). Transcriptional activity scales with genomic separation From the latent state trajectories revealed by our inference approach, we estimated the sur- vival curves of the transcriptionally active state (see the materials and methods, section 2.4, and Fig. 2C). We found a median tran- scriptional lifetime independent of genomic separation of 10 ± 5 min (SD across separa- tions; Fig. 2D). This corresponds to about three to five independent rounds of transcription on average, given the typical promoter switch- ing correlation time of the system (50). Sim- ilarly, the relative proportion of transcriptionally active states within the paired subpopulation is insensitive to genomic separation (Fig. 2E). Brückner et al., Science 380, 1357–1362 (2023) 30 June 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E 1 1 Fig. 2. Scaling of interlocus distances and transcriptional activity across genomic separations. (A) Probability distributions of the interlocus distances R. Distributions are separated by state, with paired states pooled across genomic separations, and individual distributions are shown for the open state. (B) Average interlocus distances hRi for each of the three transcriptional states. Blue dashed line indicates a linear best fit to the Ooff data for the range of genomic separations 58 to 190 kb, with exponent 1/d = 0.31 ± 0.07. Dashed cyan and red lines are average values of the interlocus distances of the Poff and Pon states, respectively, with shaded areas indicating SEM. Solid dark green and red lines indicate predictions for ideal and crumpled polymers, respectively. (C) Survival curves S(t) of the transcriptionally active state Pon, giving the probability that transcription remains active after time t. Orange curve: data for no-homie constructs (s = 58 kb). Curves were estimated using the Kaplan-Meier estimator, which accounts for censoring that occurs if the trajectory begins or ends in the transcriptionally active state (81). Shaded areas show 95% confidence intervals (see the materials and methods, section 2.4). (D) Median lifetime of the transcriptionally active state Pon as a function of genomic separation using the Kaplan-Meier estimator (dots) and a maximum-likelihood estimator assuming exponential decay of the survival curves (triangles) (see the materials and methods, section 2.4). (E) Probability of the paired on and off states conditioned on the system being in one of these two paired configurations. (F) Overall probability of the paired configurations Poff and Pon as a function of genomic separation. Gray line is the best fit with exponent 0.9 ± 0.2. Green and dark red lines indicate predicted exponents for the contact probabilities of the ideal and crumpled chain polymer models, respectively. By contrast, the overall probability of observ- ing either of the paired configurations decreases with genomic separation and exhibits a power- law scaling P(s) ~ s–f, with f = 0.9 ± 0.2 (Fig. 2F). Because transcriptional lifetimes are in- dependent of distance, the scaling of P(s) is likely dominated by the search of the two loci to come into contact. Different polymer mod- els make distinct predictions of the scaling of contact probabilities (22, 33, 51). For ideal chains, f = 3/2, whereas crumpled chains ex- hibit f ≈ 1.15 (52), which is close to the scaling that we observed. To determine how these results depend on the nature of the homie insulator–mediated focal contacts in our system, we used a re- porter construct in which the homie sequence was replaced by a l DNA sequence of the same length. At a 58-kb separation, transcriptional encounters still occur, albeit with a shorter median lifetime of 4.9 ± 1.2 min (Fig. 2C and fig. S13). Furthermore, the probability of observ- ing a transcriptional state was reduced from (30 ± 5)% for the homie version to (8.5 ± 0.8)% in the no-homie version. By contrast, very few such encounters were found for a 149-kb no- homie separation (8), where contact probabil- ity decreases from (6 ± 1)% to >1% when the homie sequence was replaced by a l DNA. Together, these results demonstrate quanti- tatively how both genomic sequence and geno- mic separation control the rate of transcriptional encounters. The scaling of transcription prob- abilities with separation suggests that the transition from the open to the paired config- uration is a key limiting step in transcriptional activation of distal DNA loci, which is limited by the time taken to diffuse into proximity. Characterizing the subdiffusive locus search process To understand these diffusive timescales, we considered the real-time dynamics of the blue- and green-labeled DNA loci. We found that the majority of single-cell trajectories sampled the whole range of physical distances in each topological state, because they showed a simi- lar spread as the ensemble-averaged distri- bution (Fig. 3, A to C, and fig. S8). Thus, rather than existing in constrained configurations as observed in other genomic contexts (41), this observation supports the picture of a dy- namic search process exploring a broad range of distances. Þ (cid:2) ri t0ð ÞÞ2it0 We quantified how this search process is reflected in the motion of individual DNA loci by computing the single-locus MSD M1 tð Þ ¼ hðri t0 þ t ¼ Gtb , where ri(t) is ð the 3D position of the locus, G is diffusivity, and b is the dynamic exponent. This expo- nent quantifies how locus diffusion scales with time and can be related theoretically to the packing of the chromosome through the frac- tal dimension d: b = 2/(2 + d) (37, 39, 53). Although the ideal chain model predicts b = 1/2 (54), we expected b = 2/5 for a crumpled polymer (37). Our system showed a scaling exponent of b = 0.52 ± 0.04 across genomic separations (error bar: SD calculated from Brückner et al., Science 380, 1357–1362 (2023) 30 June 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E total variance across separations) for both the endogenous eve locus (blue) and the ectopic reporter (green), which is close to the predic- tion of the ideal chain model and consistent with previous works (26, 27, 35) (see the mate- rials and methods, section 3, and Fig. 3, D and E). Our data further indicate that the single- locus dynamics are not affected by transcrip- tional activity, unlike previous accounts (43), because they were consistent across the three topological states (Fig. 3F). ð To further understand how the locus dynamics are determined by the interplay of chromosome organization and single-locus dynamics, we an- alyzed the joint dynamics of the two coupled chromosomal loci. From the statistics of the 3D distance vector R(t), we computed the two- Þ (cid:2) R t0ð ÞÞ2it0 locus MSD M2 tð Þ ¼ hðR t0 þ t (26), which quantifies the crossover between two intuitive regimes. Whereas at small time lags, the MSD is determined by the indepen- dent diffusion of the two loci [M2(t) = 2Gtb], it exhibits a crossover to a plateau at large times, given by the average squared interlo- cus distance [M2(t) = 2hR2i] (see the ma- terials and methods, section 5, and Fig. 4, A and B). Consistent with the observed single- locus dynamics, we found that the subdiffu- sive regime of the experimental two-locus MSD exhibited an exponent close to 1/2 for those datasets in which this regime was sam- pled (Fig. 4A and fig. S16). Similarly, for large time lags, the two-locus autocorrelation re- vealed agreement with the ideal chain scal- ing (Fig. 4, C and D). Thus, the full time dependence of the MSD is well described by the ideal chain predictions, both for single and coupled loci. Interlocus relaxation times exhibit an anomalous scaling with genomic separation Having established the static and dynamic properties of the system, we next investigated the consequences of these features for the time- scales of the two-locus search process. This process is determined by the interplay of chro- mosome dynamics and organization and can be characterized by a relaxation time t, which corresponds to the timescale of the crossover of the two regimes of the two-locus MSD (Fig. 2A). Specifically, t is the time taken by the two loci to diffuse (dynamics) over their typical dis- tance of separation (organization): Gtb ~ s2/d. This relationship predicts a scaling of relax- ation times with genomic separation t ~ sg. For ideal chains with fractal dimension d = 2 and a diffusion exponent b = 1/2, this yields the classical result g = 2. By contrast, for crum- pled chains, b = 2/5 and d = 3, yielding g = 5/3 (see the materials and methods, section 5, and table S7). To infer the relaxation time in our data as a function of genomic separation, we performed a Bayesian fitting of the two-locus MSD with Fig. 3. Dynamics of DNA locus search and single-locus fluctuations. (A) Single-cell interlocus distance trajectories for the three topological states (s = 149 kb). For each state, 80 trajectories are shown, with one sample trajectory highlighted in bold. (B) Distance distributions (bar histogram) of the highlighted trajectory in (C) compared with the ensemble distribution obtained by averaging over all cells (line). (C) Single-cell interlocus distance distributions (thin lines) of all trajectories for the three states compared with ensemble distributions in bold (s = 149 kb). Distributions are smoothed using Gaussian kernel density estimation with a width of 100 nm. Only trajectories with at least 10 time points are included to ensure sufficient statistics for comparison. (D) Single-locus MSDs for all genomic separations (color code corresponds to Fig. 2A). Single-locus MSDs were calculated by estimating 3D MSDs from motion-corrected trajectories in the x-y plane of the system (see the materials and methods, section 3). Open data points correspond to a shorter imaging time interval Dt = 5.4 s (s = 149 kb). (E) Single-locus MSDs comparing enhancer (blue) and promoter (green) fluctuations (s = 149 kb). (F) Single-locus MSDs comparing fluctuations in the three states (s = 149 kb). the ideal chain expression (26) (see the mate- rials and methods, section 4.1). We found that the fitted two-locus diffusion coefficient in- creased with genomic separation up to 595 kb, with an approximate scaling G(s) ~ s0.27 ± 0.03 (Fig. 4E). This scaling appeared to plateau for the largest genomic separation (3.3 Mb) at a value close to the single-locus diffusion, which remained approximately constant across separa- tions (Fig. 4E). The absolute value of the diffu- sivity at the plateau was almost 20-fold larger than previous measurements in mammalian stem cells with similar genomic separation (26), suggesting comparatively fast chromosome dy- namics (fig. S23). The relaxation time was determined by com- bining our estimate of the two-locus diffusivity with the average interlocus distances. The com- bination of static and dynamic exponents in our system, as well as the scale-dependent dif- fusivity, results in an anomalous scaling of relaxation times with genomic separation with an exponent g = 0.7 ± 0.2 (Fig. 4F). This ex- ponent corresponds to a much shallower scal- ing with separation than predicted by either the ideal or crumpled chain theory. This result was further confirmed by a data collapse of the two-locus autocorrelation functions (Fig. 4D and fig. S20). Although these results are de- rived from the trajectories in the Ooff state, they are insensitive to the details of the state inference (fig. S11). In sum, the key result here is that the relaxation time, which sets the time- scale of two-locus encounters, is much less de- pendent on genomic separation than predicted by existing polymer models. Anomalous relaxation time scaling induces long-ranged velocity correlations ð Þit0 t0 þ t tð Þ ¼ hv dð Þ The anomalous relaxation time scaling makes a key prediction for the correlations of the ab- solute motion of DNA loci, quantified by the velocity cross-correlation C dð Þ t0ð Þ(cid:3) vv v dð Þ . These correlations are deter- j mined by the relaxation time through the dimensionless ratio d/t, where d is the exper- imental observation timescale (Fig. 4G) (55). Having determined the relaxation times t, one can therefore make a parameter-free prediction of the correlations, which decay i Brückner et al., Science 380, 1357–1362 (2023) 30 June 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E 1 1 ð Þi t0 Fig. 4. Joint dynamics of DNA locus pairs. (A) Ideal chain Rouse prediction of the two-locus MSD M2(t) = 2Gt1/2(1 – e–t/pt) + 2J erfc[(t/pt)1/2] (26) (gray line) using best-fit values G, J, b = 1/2, and t = (J/G)2, compared with experiment (s = 595 kb). Green and red lines give expected scaling tb for t ≪ t for the generalized Rouse model for ideal and crumpled chains, respectively (see the materials and methods, section 5). (B) All experimental two-locus MSDs with relaxation times (dashed lines) and expected asymptotes 2hR2i (solid lines; color code corresponds to Fig. 2A). (C) Scaling of the diffusion coefficients G from two-locus MSD fits (black dots) compared with single-locus diffusion coefficients obtained from single-locus MSDs (Fig. 3, F to H). Dashed line is the best fit to two-locus diffusivity with exponent 0.27 ± 0.03 (s = 58 to 595 kb); solid lines are the average value of single- locus diffusivities; shaded area shows SE calculated from total variance across separations. (D) Two-locus auto- correlation function (ACF) C2 tð Þ ¼ hR t0ð Þ(cid:3) R t0 þ t ¼ hR2i (cid:2) M2 tð Þ=2 (gray) compared with data (s = 149 kb). Green and red curves indicate the power-law exponent l = 2(1 – d)/(2 + d) of the correlation function C2(t) ~ tl for ideal and crumpled chains for t ≫ t, respectively (39). (E) Collapsed correlations C2 ~ C2(ts–g)/hR2i with g = 0.7. Inset: raw correlations C2(t) for varying genomic separation. Open data points correspond to data obtained with a higher sampling rate. (F) Scaling of inferred relaxation times compared with predicted ideal and crumpled chain exponents. Gray line is the best fit with exponent g = 0.7 ± 0.2. (G) Predicted velocity cross-correlation functions C dð Þ hv dð Þ Þi t0 dimensionless ratio d/t (55). Velocities are calculated on a time interval d as v(d)(t) = [x(t + d) – x(t)]/d. (H) Scaling of the zero-time velocity cross-correlation intercept normalized by the zero-time autocorrelation, vv 0ð Þ=C dð Þ C dð Þ v 0ð Þ, for the Ooff (blue) and Pon (red) states; d = 300 s. Green line is the prediction based on ideal chain Rouse scaling of the relaxation times (g = 2) with an intercept determined based on the 58-kb data point; gray line is the parameter-free prediction using the inferred anomalous relaxation time scaling (g ≈ 0.7) (see the materials and methods, section 4.3); dashed red line is the average correlation in the Pon state. vv for increasing values of the t0ð Þ (cid:3) v dð Þ ð t0 þ t tð Þ ¼ i j substantially more slowly than for the ideal Rouse model (see the materials and methods, section 4.3, and Fig. 4H, green and gray lines). We found that the experimental correlations were quantitatively captured by this parameter- free prediction (Fig. 4H), including the full time dependence of the correlations (fig. S22). This demonstrates that the anomalous relaxa- tion time scaling indeed leads to long-range velocity cross-correlations of chromosomal loci, pointing toward potential long-range interactions. Discussion We developed an experimental approach to perform in vivo imaging of the joint dynam- ics of enhancer-promoter pairs with varying genomic separation and simultaneous moni- toring of their transcriptional output. Observ- ing the dynamics of pairs of DNA loci has only become possible recently and has been done for tagged DNA loci at a single fixed genomic separation (8, 26, 27, 36). Here, we show how imaging across genomic separations gives insight into the relative motion, dynamic encounters, and transcriptional activation of such loci. Many features of the two-locus dynamics, including the subdiffusive exponent close to 0.5, are very well conserved with measure- ments of CTCF sites at TAD boundaries in mammalian systems (26, 27), despite CTCF not being essential for Drosophila embryo- genesis (56). In absolute numbers, however, our measurements revealed large diffusion coefficients of DNA loci that are ~20-fold larger than in mammalian cells (26) (fig. S23). Early fly development follows a tight sched- ule, suggesting that the chromosome dynam- ics may have evolved to operate on much faster timescales than mammalian systems. By contrast, the median lifetime of focal con- tacts in our system of 12 ± 5 min is well within the range of typical CTCF loop lifetimes of 10 to 30 min in mammalian cells (26, 27). These timescales facilitate transcriptional lifetimes of 10 ± 5 min in our system, which in the ab- sence of the homie insulator are reduced to 4.9 ± 1.2 min (Fig. 2C and fig. S13), highlighting the importance of focal elements for contact formation in Drosophila. Brückner et al., Science 380, 1357–1362 (2023) 30 June 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E To initiate such transcriptional encoun- ters, the two loci perform a search process to reach physical proximity. The timescale of this search process is given by the lifetimes of the unpaired Ooff state, which depends on multi- ple factors. These factors typically include the landscape of the search process (57, 58), the biochemical binding properties of the focal elements when proximity has been reached (59), and the correlation time of the system (i.e., the relaxation time). Indeed, we found that lifetimes increased with the relaxation times and were ~10 times larger on average (fig. S19). We have demonstrated how key features of our system, tight crumpled chain packing, sub- diffusion with exponent 0.5, and a separation- dependent two-locus diffusivity, lead to relaxation times that are much less dependent on genomic separation than predicted by existing polymer models. Indeed, for an ideal Rouse polymer, the relaxation time for our largest genomic separation (3.3 Mb) would be ~3000 times longer than for the shortest 58-kb separation. Our measurements, however, revealed that it only takes ~20 times longer, corresponding to a >100-fold reduction. This reduced depen- dence on distance implies that transcriptional encounters are possible across large genomic distances, allowing enhancers dispersed across the chromosome to find their target promoter efficiently. This might be one of the reasons that evolution can act on distal sequences from a given target promoter. Overall, our find- ings have crucial implications for the spatio- temporal organization of the cell nucleus, including the dynamics of long-range focal con- tacts (28) and mammalian enhancer-promoter interactions (9–12, 44). From a polymer physics perspective, our measured exponents suggest that the rela- tionship between static and dynamic prop- erties in the generalized Rouse framework, which relies on the assumption of local fric- tion, does not apply to chromosomes. This implies that long-range interactions such as hydrodynamics or active motor-mediated interactions (60, 61) could play a role. Indeed, the simplest polymer model that relaxes the Rouse assumption and includes long-range hydrodynamic interactions, the Zimm mod- el (54), predicts a scaling relationship of re- laxation times with genomic separations with an exponent of g = 1 (table S7), which is close to our measured value of g ≈ 0.7. Furthermore, the observed separation-dependent diffusivity points to additional interactions or heteroge- neities along the polymer. Such heterogeneities could be caused by a number of processes, such as cross-linking (41), out-of-equilibrium activity (61), entanglements (62), or the pres- ence of condensates (18–20). Together, these processes may orchestrate the anomalous scaling of relaxation times with genomic sep- aration. In future work, the mechanistic un- derpinnings of our findings should be tested using polymer simulations (40, 41, 51, 63–69) to generate hypotheses for new sets of experiments. RE FERENCES AND NOTES 1. W. Flemming. Zellsubstanz, Kern Und Zelltheilung (Vogel, 1882). 2. R. Milo, R. Phillips, Cell Biology by the Numbers (CRC Press, 2015). 3. R. Stadhouders, G. J. Filion, T. Graf, Nature 569, 345–354 (2019). 4. B. van Steensel, E. E. M. Furlong, Nat. Rev. Mol. Cell Biol. 20, 327–337 (2019). 5. M. Levine, Curr. Biol. 20, R754–R763 (2010). 6. T. K. Kim, R. Shiekhattar, Cell 162, 948–959 (2015). 7. D. Vernimmen, W. A. Bickmore, Trends Genet. 31, 696–708 (2015). 8. H. Chen et al., Nat. Genet. 50, 1296–1303 (2018). 9. B. Tolhuis, R. J. Palstra, E. Splinter, F. Grosveld, W. de Laat, Mol. Cell 10, 1453–1465 (2002). 10. V. V. Uslu et al., Nat. Genet. 46, 753–758 (2014). 11. Y. Zhang et al., Nature 504, 306–310 (2013). 12. A. Sanyal, B. R. Lajoie, G. Jain, J. Dekker, Nature 489, 109–113 (2012). 50. B. Zoller, S. C. Little, T. Gregor, Cell 175, 835–847.e25 (2018). 51. A. Rosa, N. B. Becker, R. Everaers, Biophys. J. 98, 2410–2419 (2010). 52. J. D. Halverson, W. B. Lee, G. S. Grest, A. Y. Grosberg, K. Kremer, J. Chem. Phys. 134, 204904 (2011). 53. P. G. De Gennes, Macromolecules 9, 587–593 (1976). 54. M. Rubinstein, R. H. Colby, Polymer Physics (Oxford Univ. Press, 2003). 55. T. J. Lampo, A. S. Kennard, A. J. Spakowitz, Biophys. J. 110, 338–347 (2016). 56. A. Kaushal et al., Nat. Commun. 12, 1011 (2021). 57. G. Wilemski, M. Fixman, J. Chem. Phys. 60, 866–877 (1974). 58. A. Szabo, K. Schulten, Z. Schulten, J. Chem. Phys. 72, 4350–4357 (1980). 59. M. Fujioka, H. Mistry, P. Schedl, J. B. Jaynes, PLOS Genet. 12, e1005889 (2016). 60. R. Bruinsma, A. Y. Grosberg, Y. Rabin, A. Zidovska, Biophys. J. 106, 1871–1881 (2014). 61. M. M. Tortora, H. Salari, D. Jost, Curr. Opin. Genet. Dev. 61, 37–43 (2020). 62. R. Kawamura et al., J. Cell Biol. 188, 653–663 (2010). 63. A. Rosa, R. Everaers, PLOS Comput. Biol. 4, e1000153 (2008). 64. M. Barbieri et al., Proc. Natl. Acad. Sci. U.S.A. 109, 16173–16178 (2012). 65. L. Giorgetti et al., Cell 157, 950–963 (2014). 66. L. Liu, G. Shi, D. Thirumalai, C. Hyeon, PLOS Comput. Biol. 14, 13. Y. Huang, R. Neijts, W. de Laat, FEBS Lett. 594, 3439–3449 e1006617 (2018). (2020). 14. J. R. Dixon et al., Nature 485, 376–380 (2012). 15. C. Hou, L. Li, Z. S. Qin, V. G. Corces, Mol. Cell 48, 471–484 (2012). 16. E. P. Nora et al., Nature 485, 381–385 (2012). 17. T. Sexton et al., Cell 148, 458–472 (2012). 18. A. Boija et al., Cell 175, 1842–1855.e16 (2018). 19. W.-K. Cho et al., Science 361, 412–415 (2018). 20. M. Conte et al., Nat. Commun. 13, 4070 (2022). 21. C. S. Osborne et al., Nat. Genet. 36, 1065–1071 (2004). 22. E. Lieberman-Aiden et al., Science 326, 289–293 (2009). 23. O. Symmons et al., Genome Res. 24, 390–400 (2014). 24. T. J. Stevens et al., Nature 544, 59–64 (2017). 25. M. V. Arrastia et al., Nat. Biotechnol. 40, 64–73 (2022). 26. M. Gabriele et al., Science 376, 496–501 (2022). 27. P. Mach et al., Nat. Genet. 54, 1907–1918 (2022). 28. M. Levo et al., Nature 605, 754–760 (2022). 29. S. Wang et al., Science 353, 598–602 (2016). 30. B. Bintu et al., Science 362, eaau1783 (2018). 31. A. Grosberg, Y. Rabin, S. Havlin, A. Neer, Europhys. Lett. 23, 373–378 (1993). 32. A. Yu. Grosberg, S. K. Nechaev, E. I. Shakhnovich, J. Phys. (Paris) 49, 2095–2100 (1988). 33. L. A. Mirny, Chromosome Res. 19, 37–51 (2011). 34. S. C. Weber, A. J. Spakowitz, J. A. Theriot, Phys. Rev. Lett. 104, 238102 (2010). 35. J. S. Lucas, Y. Zhang, O. K. Dudko, C. Murre, Cell 158, 339–352 (2014). 36. J. M. Alexander et al., eLife 8, e41769 (2019). 37. M. V. Tamm, L. I. Nazarov, A. A. Gavrilov, A. V. Chertovich, Phys. Rev. Lett. 114, 178102 (2015). 38. S. Shinkai, T. Nozaki, K. Maeshima, Y. Togashi, PLOS Comput. Biol. 12, e1005136 (2016). 39. K. E. Polovnikov, M. Gherardi, M. Cosentino-Lagomarsino, M. V. Tamm, Phys. Rev. Lett. 120, 088101 (2018). 40. M. Nicodemi, A. Pombo, Curr. Opin. Cell Biol. 28, 90–95 (2014). 41. N. Khanna, Y. Zhang, J. S. Lucas, O. K. Dudko, C. Murre, Nat. Commun. 10, 2771 (2019). 42. T. Germier et al., Biophys. J. 113, 1383–1394 (2017). 43. B. Gu et al., Science 359, 1050–1055 (2018). 44. J. Zuin et al., Nature 604, 571–577 (2022). 45. S. V. Ulianov et al., Nat. Commun. 12, 41 (2021). 46. B. L. Hua, T. L. Orr-Weaver, Genetics 207, 29–47 (2017). 47. B. Avşaroğlu et al., PLOS ONE 9, e102474 (2014). 48. J. Dekker, M. A. Marti-Renom, L. A. Mirny, Nat. Rev. Genet. 14, 390–403 (2013). 49. L. Barinov, S. Ryabichko, W. Bialek, T. Gregor, Transcription- 67. A. Buckle, C. A. Brackley, S. Boyle, D. Marenduzzo, N. Gilbert, Mol. Cell 72, 786–797.e11 (2018). 68. C. A. Brackey, D. Marenduzzo, N. Gilbert, Nat. Methods 17, 767–775 (2020). 69. H. Salari, M. Di Stefano, D. Jost, Genome Res. 32, 28–43 (2022). 70. D. B. Brückner, B. Zoller, dbrueckner/DynamicChromosome: initial release, Zenodo (2023); https://zenodo.org/record/ 7926778. 71. D. B. Brückner, H. Chen, L. Barinov, B. Zoller, T. Gregor, Trajectory data for: Stochastic motion and transcriptional dynamics of pairs of distal DNA loci on a compacted chromosome, Zenodo (2023); https://zenodo.org/record/ 7616203. AC KNOWLED GME NTS We thank R. Evearers, L. Giorgetti, A. Grosberg, S. Grosse-Holz, M. Levo, L. Mirny, A. Rosa, V. Scolari, A. Spakowitz, and G. Tkacik for helpful comments and discussion; K. Bystricky for introducing us to the ParB/parS system; and F. Payre and P. Valenti for sharing the ParB-eGFP plasmid and the parS sequence. Funding: This work was supported in part by the U.S. National Science Foundation, the Center for the Physics of Biological Function (grant PHY-1734030), and the National Institutes of Health (grants R01GM097275, U01DA047730, and U01DK127429). D.B.B. was supported by the NOMIS Foundation as a fellow and by an EMBO postdoctoral fellowship (ALTF 343-2022). H.C. was supported by a Charles H. Revson Biomedical Science Fellowship. Author contributions: H.C. and T.G. designed the experiments. H.C. and L.B. performed the experiments and analyzed the images. D.B.B. and B.Z. analyzed the data. D.B.B. and T.G. interpreted the data. D.B.B., B.Z., and T.G. wrote the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: The code and software used in this study (70) and the raw trajectory data (71) are freely available through Zenodo. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf5568 Materials and Methods Figs. S1 to S23 Tables S1 to S8 References (72–86) Movies S1 to S7 MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. dependent spatial organization of a gene locus. arXiv:2012. 15819 [q-bio.MN] (2020). Submitted 29 October 2022; accepted 31 May 2023 10.1126/science.adf5568 Brückner et al., Science 380, 1357–1362 (2023) 30 June 2023 6 of 6
10.1126_science.adg0144
RES EARCH SENSATION PIEZO2 and perineal mechanosensation are essential for sexual function Ruby M. Lam1,2, Lars J. von Buchholtz3, Melanie Falgairolle1, Jennifer Osborne1, Eleni Frangos1, M. Rocio Servin-Vences4, Maximilian Nagel1, Minh Q. Nguyen3, Monessha Jayabalan1, Dimah Saade5, Ardem Patapoutian4, Carsten G. Bönnemann5, Nicholas J. P. Ryba3, Alexander T. Chesler1,5* Despite the potential importance of genital mechanosensation for sexual reproduction, little is known about how perineal touch influences mating. We explored how mechanosensation affords exquisite awareness of the genitals and controls reproduction in mice and humans. Using genetic strategies and in vivo functional imaging, we demonstrated that the mechanosensitive ion channel PIEZO2 (piezo-type mechanosensitive ion channel component 2) is necessary for behavioral sensitivity to perineal touch. PIEZO2 function is needed for triggering a touch-evoked erection reflex and successful mating in both male and female mice. Humans with complete loss of PIEZO2 function have genital hyposensitivity and experience no direct pleasure from gentle touch or vibration. Together, our results help explain how perineal mechanoreceptors detect the gentlest of stimuli and trigger physiologically important sexual responses, thus providing a platform for exploring the sensory basis of sexual pleasure and its relationship to affective touch. S exual reproduction is a fundamental driver for animal behavior, and adap- tations required for courtship, including sexual ornamentation and ritual dis- plays, are cornerstones of evolutionary theory (1–3). Visual, auditory, and olfactory cues promote mating in various mammalian species (4–8); however, the act of copulation itself can be considered a specialized form of touch endowed with its own cortical field (9). Although the discovery of the mechanically gated ion-channel PIEZO2 (piezo-type mecha- nosensitive ion channel component 2) (10) has spurred remarkable progress in our under- standing of discriminative touch (11, 12), far less is known about mechanosensation in the genitals (13, 14), including how it triggers physiological responses and elicits pleasure. We hypothesized that sexual touch might exhibit unusual response specialization to con- trol mating and provide affective and mo- tivational feedback. We also expected that there would be sexual dimorphism both in sensation and in responses triggered by genital- innervating mechanosensors. To test these hypotheses, we developed a series of behav- ioral and functional imaging assays to probe the role of PIEZO2 in genital mechanosen- sation and sexual function. In addition, by exploring the impact of PIEZO2 loss of func- tion caused by a rare inherited syndrome, we 1National Center for Complementary and Integrative Health (NCCIH), Bethesda, MD 20892, USA. 2Brown-National Institutes of Health Graduate Partnerships Program, Brown University, Providence, RI 02912, USA. 3National Institute of Dental and Craniofacial Research, Bethesda, MD 20892, USA. 4Howard Hughes Medical Institute, Department of Neuroscience, Dorris Neuroscience Center, The Scripps Research Institute, La Jolla, CA 92037, USA. 5National Institute of Neurological Disorders and Stroke, Bethesda, MD 20892, USA. *Corresponding author. Email: alexander.chesler@nih.gov determined how these findings relate to hu- man sexual experience. Unusual sensitivity and PIEZO2 dependence of perineal touch A standard touch sensitivity test uses cal- ibrated von Frey filaments to measure detec- tion threshold. In mice, von Frey sensitivity of the glabrous hind-paw and hairy skin of the face have similar withdrawal thresholds (15–17) despite very different patterns of innervation (18). We adapted this assay to compare stim- ulation of the perineum (the region extend- ing from the anus to the genitals in male and female mice) with that of the plantar surface of the paw. In the hind-paw assay, mice res- pond by withdrawing the paw with no indi- cation of pain or distress. Our data (Fig. 1A) match literature reports, with filaments ≥0.4 g eliciting responses in the majority of trials, but filaments ≤0.16 g rarely provoking reaction (15, 17). By contrast, stimulation of the peri- neum evoked a highly stereotyped startle and investigative response (movie S1) both in male and female mice. Even the finest filament avail- able (0.008 g) elicited this reaction from every animal (Fig. 1B), demonstrating exquisite sensi- tivity of the perineum to forces below those that reliably trigger responses from other sites, even in mice with profound allodynia (15, 17); female mice were marginally but consistently more sensitive than males (Fig. 1B). The mechanically activated ion channel PIEZO2 is essential for discriminative touch in mice and humans (11, 12). We anticipated that this mechanoreceptor would also be re- sponsible for the sensitivity of the perineum. Piezo2-null mice die as neonates (19); therefore, we generated conditional genetic deletions using a Hoxb8-Cre line (Piezo2Hoxb8) to target cells below the mid-thoracic region (17). We used this strategy to assess the role of PIEZO2 in perigenital sensation and observed profound loss of behavioral response to von Frey fila- ments (Fig. 1C), with the highest force tested (1.4 g) eliciting responses in only ~40% of trials (movie S1). Local inhibition of the perineum with lidocaine attenuated von Frey responses of controls (fig. S1), and Piezo2Hoxb8 responses to noxious mechanical pinprick were indistinguish- able from those of controls (Fig. 1C and movie S1). Therefore, the Piezo2Hoxb8 deficit is likely to be sensory rather than related to a move- ment disorder (20). These experiments dem- onstrated that PIEZO2 is crucial for triggering behavioral responses to the gentlest of peri- genital touch in mice; without this touch re- ceptor, von Frey stimulation of the genital region rarely elicited responses even at inten- sities considered noxious. We previously studied a rare cohort of peo- ple with biallelic loss-of-function variants of PIEZO2 who have sensory deficits fully con- sistent with those described in animal models (11, 21). In our clinical interviews, five adult hu- man subjects with PIEZO2-deficiency syndrome (three male and two female) reported severe hyposensitivity in genital sensation (table S1); however, comprehensive quantitative testing has not been possible. One individual adult male consented to quantitative sensory testing of his genitalia during clinical evaluation. His penile von Frey detection threshold (3.1 ± 1.5 g) was far higher than values reported in the litera- ture: 0.3 to 0.6 g in a similar location (22). He had difficulty detecting pressure below 1 kg/cm2 at the midshaft and was insensitive to strong vibration at 50 and 100 Hz, which is consistent with our findings in mice. By contrast, litera- ture values for penile fine-touch pressure thresholds in a range of healthy men are far lower (23), and vibration is normally readily detected (23). Anatomy of perineal neurons Somatosensory neurons in the lower body have soma in lumbar (L1 to L6) and sacral (S1 to S4) dorsal root ganglia (DRG) (24). However, few details about the types or sensitivity of neurons that target the genitals are known. Multicolor cholera toxin subunit-b (CTB) tracing from both hind-paw and genitals robustly labeled neurons in S1 and S2 DRG (fig. S2A) and distinguished neurons that target perigenital subregions (Fig. 1, D and E). Injections to the perineum, prepuce, and glans (male mice) or vaginal open- ing (female mice) resulted in largely nonover- lapping labeling of neurons with a range of cell diameters (Fig. 1D and fig. S2B). In both sexes, dense projections targeted the (L6 to S2) spinal cord, with perineal neurons (Fig. 1, D and E, cyan) synapsing in the touch recip- ient zone (25) of the lateral dorsal horn (Fig. 1E and fig. S2C). Neurons innervating male pre- puce (Fig. 1, D and E, yellow) projected to a Lam et al., Science 381, 906–910 (2023) 25 August 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E medial portion of the touch zone (Fig. 1E), whereas glans axons (Fig. 1, D and E, magenta) terminated proximal to the central canal (Fig. 1E). In females, axons from the vaginal opening targeted the medial dorsal horn, whereas those from the prepuce, which includes the clitoris, closely resembled those from the glans in males. To visualize the peripheral anatomy of touch neurons in the perineum, we generated mice in which Piezo2-expressing neurons were se- lectively labeled by crossing a Piezo2-Cre allele (26) into a neural-specific Snap25-LSL-GFP re- porter line (27). Green fluorescent protein (GFP) staining of cleared skin demonstrated that the perineum was densely innervated with lanceolate and circumferential endings surrounding hair follicles (Fig. 1F), which is consistent with innervation by a broad range of low-threshold mechanosensory neurons (LTMRs) and the PIEZO2-dependent behav- ioral sensitivity of mice to perineal touch. Perineal sensory neurons exhibit high sensitivity to punctate stimulation We developed a sacral ganglia imaging prep- aration to compare neural responses to a range of gentle and intense mechanical stimuli (28) applied to the hind-paw and perineum (fig. S3 and movie S2). Neurons innervating paw gla- brous skin divided into LTMRs and high- threshold mechanosensory neurons (HTMRs) on the basis of their response selectivity (Fig. 2A; fig. S3; and supplementary materials, mate- rials and methods). HTMRs that innervate the paw outnumbered LTMRs by a factor of 2. In particular, LTMRs exhibited graded von Frey sensitivity (Fig. 2A and fig. S3D) and could be activated by forces as low as 0.008 g, whereas HTMRs were essentially silent at forces below 0.4 g, matching behavioral withdrawal thresh- old and implicating HTMRs in this response. By contrast, perineal sensation was dominated by LTMRs, with ~60% of mechanosensory neu- rons responding to gentle stimuli (Fig. 2B and figs. S3 and S4) and broad similarity between male and female mice (fig. S4). Almost all per- ineal mechanosensors could be activated by von Frey stimulation (Fig. 2B, figs. S3, S4), and their calcium (GCaMP) signals were markedly stronger than for paw-innervating neurons (fig. S3D). Few HTMRs responded to the fine filaments that reliably evoked behavioral re- sponses (Fig. 1B and movie S1). Therefore, both male and female mice are attuned to perineal LTMR input, and the stereotyped reaction to genital touch is not a sign of pain. A broad role for PIEZO2 in perineal sensation To measure the contribution of PIEZO2 to perigenital touch and to dissect the mecha- nism underlying the extreme sensitivity to von Frey stimulation, we next used the sacral imaging platform to selectively image cells that lack this stretch-gated ion channel (fig. S4D). Fig. 1. Behavioral sensitivity of mice to perineal touch and underlying anatomy. (A to C) Reaction of mice to punctate touch (A) wild-type hind-paw, (B) wild-type perineum, and (C) Piezo2Hoxb8 perineum. (Left) Example responses for individual mice (points and thin lines; four males and four females) and mean (solid lines) to a series of calibrated von Frey filaments (grams, each tested 10 times per mouse). (Middle) Quantitation of von Frey threshold (≥5/10; n = 12 males and 12 females). Thresholds are different between all three groups [one-way analysis of variance (ANOVA) on ranks P < 0.001]. Wild-type females exhibited a lower perineal touch threshold than that of males (Mann-Whitney t test; P < 0.0001); there were no significant differences in other responses (supplementary materials, statistical reporting). (D and E) Triple-color retrograde CTB tracing from the perineum (cyan), prepuce (yellow), and glans (magenta) showing (D) cell bodies of lumbar-sacral sensory neurons in the DRG and (E) termini in the dorsal spinal cord. The dotted line indicates approximate extent of dorsal horn. In (E) and (F), n = 4 mice. Scale bars, 100 mm. (F) Anatomy of sensory ending of Piezo2-expressing sensory neurons in the perineum. (Inset) A magnified view of a single hair (boxed) highlighting prominent lanceolate and circumferential endings (n = 2 males and 1 female). Scale bar, 50 mm. As expected, deletion of Piezo2 (Piezo2cKO) dra- matically affected the mechanosensitivity of genital-innervating neurons. The great major- ity of responses to air puff, vibration, and brush were eliminated. Thus, mechanosensory neu- rons were only stimulated by pinch and were almost exclusively HTMRs (fig. S4E). The overall number of HTMRs was similar between wild- type and Piezo2cKO mice (fig. S4F), which is consistent with earlier studies (17, 20, 21). Piezo2cKO mice responses to von Frey stim- ulation were substantially reduced and reca- pitulated those of control perineal HTMRs (Fig. 2, C to E). These results likely explain the absence of behavioral reactions to von Frey stimulation in Piezo2Hoxb8 mice (Fig. 1C), sup- port the hypothesis that perineal LTMRs drive this characteristic withdrawal in wild-type mice (movie S1), and are consistent with human re- ports and sensory testing (table S1). A subset of touch neurons is required for mechanically induced erection responses Perineal investigation and touch precedes mating in many species, including mice (29). These behaviors are linked to motivational drive in both partners and trigger physiological reflexes. For example, gentle retraction of the prepuce induces penile cupping (erection) and flipping (ejaculation) in spinalized rodents (30). We reasoned that mechanosensory input drives the erection reflex and developed an assay to monitor this in restrained awake mice. A soft, transparent tube was used to gently retract the prepuce, allowing the physiological erec- tion reflex (extension of the penis into the tube) Lam et al., Science 381, 906–910 (2023) 25 August 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E A Wildtype hind-paw von Frey filament (g) Air puff Brush Vibration Pinch 0.008 0.02 0.04 0.07 0.16 0.4 0.6 1.0 1.4 s R M T L s R M T H 100 0 40 0 B Wildtype perineum von Frey filament (g) Air puff Brush Vibration Pinch 0.008 0.02 0.04 0.07 0.16 0.4 0.6 1.0 1.4 s R M T L s R M T H 100 0 40 0 C cKO Piezo2 perineum von Frey filament (g) Air puff Brush Vibration Pinch 0.008 0.02 0.04 0.07 0.16 0.4 0.6 1.0 1.4 100 0 40 0 von Frey filament (g) 0.008 0.02 0.04 0.07 0.16 0.4 0.6 1.0 1.4 E n o i l g n a g r e p s l l e C 50 25 0 Control Control HTMRs Piezo2cKO Graded filaments s R M T L s R M T H D l o r t n o C O K c 2 o z e P i 40 0 Fig. 2. Functional characterization of perineal mechanoreceptors and role of PIEZO2. (A to C) Heatmaps representing calcium (GCaMP6f) responses to (left) repetitive application of naturalistic stimuli and (right) graded von Frey stimulation. LTMRs and HTMRs are separated, and relative fluorescence changes (DF/F) are colored as indicated. Scale bar, 10 s. (A) Wild-type hind-paw, n = 4 mice. (B) Wild-type perineum, n = 4 mice. (C) Piezo2cKO perineum, n = 6 mice. Additional analysis is provided in figs. S3 and S4. (D) Spatial activity maps of control and Piezo2cKO neurons to von Frey filaments. Scale indicates response intensity. Scale bar, 100 mm. (E) Quantitation of von Frey responsive neurons in control mice (gray), Piezo2cKO mice (red), and response profile of control HTMRs (black) (mean ± SEM, n = 8 control mice, n = 6 Piezo2cKO mice). Piezo2cKO mice had fewer von Frey responsive neurons at all filament strengths (Mann Whitney U test; P < 0.0087). to be scored. Wild-type controls responded in almost every single trial (Fig. 3A); isoflurane anesthesia completely eliminated responses, and local numbing of the perineum with lido- caine greatly dampened the reflex (fig. S5). As we anticipated, Piezo2Hoxb8 mice only very rarely exhibited penile extension in response to prepuce retraction (Fig. 3A). Piezo2Hoxb8 mice exhibit broad loss of touch but also have proprioceptive (and potentially other mechanosensory) deficits (17, 28). There- fore, we examined mice with more selective Piezo2 deletions. Piezo2Pvalb mice (in which Piezo2 is inactivated by using parvalbumin- driven Cre) lack proprioceptive input but still respond to gentle touch (20). These mice had perfectly normal responses to prepuce retrac- tion (Fig. 3A) despite severe ataxia. We also generated Piezo2 deletions using an Scn10a- Cre line, which is commonly used to target a broad range of nociceptors, including HTMRs (31). Perineal HTMR responses are PIEZO2 independent (Fig. 2 and fig. S4); therefore, these mice (Piezo2Scn10a) were predicted to have normal proprioception, touch, and consequently erection reflexes. Piezo2Scn10a mice walked with normal gait, and recombination of Scn10a-Cre in sacral ganglia neurons was faithful (Fig. 3, B and C, and fig. S6A), with only a few large- diameter Scn10a-negative LTMRs labeled (fig. S6A). Nonetheless, Piezo2Scn10a mice displayed severe deficits in their erection reflex, closely recapitulating the phenotype of Piezo2Hoxb8 animals (Fig. 3A) and the effects of lidocaine (fig. S5A). Single-cell sequencing data from lumbar DRG (32) and trigeminal neurons (33) validate Scn10a as a robust marker for noci- ceptors but reveal expression in c-fiber LTMRs (cLTMRs). We used in situ hybridization (ISH) to confirm coexpression of Scn10a, the cLTMR marker Tyrosine hydroxylase (Th) (24), and Piezo2 in sacral ganglia (Fig. 3C), with only very limited recombination in other potential LTMRs (fig. S6A). Because cLTMR responses to gentle mechanical stimulation depend on Piezo2 expression (34), these data strongly sug- gest a causal role for perineal cLTMR input in triggering the erection reflex. Consistent with this hypothesis, tdTomato–positive lanceolate endings (typical of cLTMRs) surround perineal hair follicles in Scn10a-Cre, Ai9 mice (Fig. 3D). Moreover, functional imaging of perineal touch responses in Scn10a-Cre, Ai95 mice revealed that neurons responding to gentle mechanical stimuli (fig. S6, B and C) had uniform small diameters, as would be expected for cLTMRs (24, 34). Severely impaired sexual function in mice lacking PIEZO2 Loss of a touch-induced erection response in Piezo2Hoxb8 males should impair mating. In- deed, 10 pairs of mating-age Piezo2Hoxb8 males and females housed together for 6 months never produced pups, whereas wild-type (C57Bl/6) controls delivered 61 litters in this time (range, five to seven litters per pair). To assess copula- tory success more directly, we also examined the frequency of vaginal plug formation after introducing virgin females in estrus to single housed males; to eliminate bias from prior ex- perience, all mice were naïve. For C57Bl/6 mice, 7 from 10 homozygous pairings had plugs after 4 hours (Fig. 3E). By contrast, plugs were never seen for Piezo2Hoxb8 male mice when paired either with Piezo2Hoxb8 or wild-type females (Fig. 3E). As predicted from their normal erection re- flexes, Piezo2Pvalb males successfully mated with C57Bl/6 females despite severe ataxia (Fig. 3E). However, Piezo2Scn10a males failed to plug re- ceptive C57Bl/6 females, substantiating the importance of PIEZO2-dependent mechano- sensory input for male mating behavior (Fig. 3E). Although loss of erection reflexes may explain why Piezo2Hoxb8 mice fail to breed, mechanosensation probably has additional Lam et al., Science 381, 906–910 (2023) 25 August 2023 3 of 5 A D F RES EARCH | R E S E A R C H A R T I C L E roles in mating. For example, female mice have similar PIEZO2-dependent perineal mechano- sensitivity to males (fig. S4) and are even more sensitive to perigenital touch (Fig. 1); Piezo2Hoxb8 females exhibited strong mating deficits when paired with wild-type males: 9 from 10 remained unplugged after 4 hours (Fig. 3E). Ethogram analysis of female intruder assays (Fig. 3F) assess motivation by quantifying stereotyped male behaviors, including partner- grooming, anogenital chemosensory investi- gation, and mounting attempts (35, 36). We analyzed behavior for 1 hour after introduction of receptive females (supplementary materials, materials and methods). Control animals ex- hibited considerable variation in mating be- havior (Fig. 3F) but in every case (n = 10 pairs of mice) engaged in chemosensory investiga- tion and mounting attempts shortly after in- troduction of the female. Similarly, pairs of Piezo2Hoxb8 males and females (n = 10 pairs) (Fig. 3F) as well as male or female Piezo2Hoxb8 mice paired with C57Bl/6 partners (n = 10 pairs in each case) (fig. S5B) exhibited strong sexually motivated behavior, not very different from controls. However, Piezo2Hoxb8 males never achieved intromission, which was regularly observed in wild-type controls. Similarly, Piezo2Scn10a males paired with receptive C57Bl/6 females showed normal sexual motivation (n = 10 males and 10 females) (fig. S5B) but without copulatory success (Fig. 3E). Moreover, Piezo2Hoxb8 females paired with C57Bl/6 males also engaged in premating behavior, including male mounting attempts (fig. S5B), but Piezo2Hoxb8 females adopted a sit-rejection posture, pre- venting intromission (37). These data show that mechanosensation plays a substantial role in productive mating and exposes dimor- phic need for PIEZO2 and gentle touch in sexual function. Impact of PIEZO2 in human sexual experience The genital sensation of a man with complete loss of PIEZO2 function and comprehensive touch- and proprioception-related studies of individuals with PIEZO2-deficiency syndrome (11, 21) demonstrate strongly conserved roles for PIEZO2 in mammalian mechanosensation. For humans, sexual experience is not simply related to reproduction but is central to large parts of many people’s social lives and behav- ior. Information from human clinical evalua- tions (n = 5; three men and two women) (table S1) provided several consistent themes about the role of gentle touch in sex. First, these in- dividuals with biallelic loss of function (table S1A) had diagnostic clinical presentation, with loss of proprioception, absent vibration sensing, highly elevated touch threshold, and scoliosis but no cognitive difficulties, and all underwent puberty without clinically relevant problems. Second, all five people with PIEZO2 deficiency reported being sexually active and able to be 100 i s n o s u r t o r P % 50 0 Control Hoxb8 Piezo2 Scn10a Pvalb Piezo2 Piezo2 Scn10a-Cre :: Ai9 tdT-immunostaining B TdT Scn10a C Scn10a Th Piezo2 Th Homozygous mating Piezo2-KO x wildtype E + (estrus) 4 hrs Plug check 100 l s g u P % 50 0 C57Bl/6 Hoxb8 Piezo2 Scn10a Pvalb Piezo2 Piezo2 Hoxb8 Piezo2 + (estrus) 1 hr 1 hr Social interaction Anogenital investigation Mount 3 2 1 3 2 1 6 / l B 7 5 C 8 b x o H 2 o z e P i 10 20 30 40 50 60 Time (mins) Fig. 3. A role for PIEZO2-dependent perineal mechanosensation in mating. (A) Physiological responses of male mice to perineal stimulation with transparent soft tubing. Penile protrusion was scored for two sets of 10 trials. Bars indicate mean ± SEM, and points indicate individual responses. Control versus Piezo2Hoxb8 mice and Piezo2Pvalb versus Piezo2Scn10a mice were different (Mann-Whitney U test; P < 0.0001; n = 18 control mice; n = 10 Piezo2-deleted mice). (B) Representative whole-mount ISH of sacral ganglion showing faithful recombination [tdTomato (TdT); magenta] of Scn10a-Cre mouse in Scn10a (green) neurons; >90% (670 of 738) TdT cells expressed Scn10a (n = 3 ganglia). Scale bar, 50 mm. (C) Example ISH of sacral ganglion section probed for Th (magenta), Scn10a (green), and Piezo2 (cyan), illustrating expression of Scn10a and Piezo2 in cLTMRs identified with Th (n = 6 ganglia). Scale bar, 50 mm. (D) Anatomy of sensory ending of Scn10a-expressing sensory neurons in the perineum (maximum projection, full-stack). (Right) Magnified and focal views of single hairs (boxed at left), highlighting lanceolate endings of Scn10a-Cre– labeled neurons (n = 2 male mice). (E) Successful mating scored by vaginal plugs (n = 10 mice). Differences are significant for C57Bl/6 versus Piezo2Hoxb8 mice (P = 0.0031) and Piezo2Pvalb versus Piezo2Scn10a mice (P = 0.0325) (Fisher’s exact test, two-tailed). (Right) Mating success for female and male Piezo2Hoxb8 mice with C57Bl/6 partners. (F) Representative ethogram plots showing sexual motivation of three isogenic pairings of C57Bl/6 and 3 Piezo2Hoxb8 mice: social interaction (gray), anogenital investigation (pale blue), and mounting attempts (red). aroused by physical genital stimulation, erotic thoughts, or videos, reflecting motivation seen in Piezo2Hoxb8 mice (Fig. 3F). Third, individuals with PIEZO2 deficiency reported delayed, at- tenuated, or absent physiological responses to gentle genital stimulation. This included clinical diagnosis of hypo-orgasmia for the male and anorgasmia for the female participants, which again is consistent with the animal model. How- ever, the five people had strategies to compen- sate for deficits in genital sensation (table S1B). Discussion Erogenous touch conveys different meanings according to circumstance; however, many key details remain unknown. We explored how de- ficits in PIEZO2-dependent mechanosensation interfere with perigenital sensation, physiolog- ical response, copulation, and reproduction. Our results demonstrate that PIEZO2-dependent touch is required for all of these in mice. Anatomical studies have identified specialized corpuscles composed of myelinated afferents likely involved in genital sensation (38, 39). Our data strongly implicate an additional type of touch neuron, the perineal cLTMRs, as crucial drivers of sexual function. Previous studies in mice and humans suggest specialized roles for cLTMRs in conveying affective and pleasur- able touch (40, 41). Thus, it is of note that five Lam et al., Science 381, 906–910 (2023) 25 August 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E individuals without PIEZO2 function described sexual activity as satisfying and rewarding despite marked mechanosensory deficits and clinical evaluations of hypo-orgasmia (men) and anorgasmia (women). We have previously shown that for humans, other types of sensory input can compensate for deficits caused by loss of PIEZO2 function (11). For example, these individuals use vision to overcome pro- prioceptive deficits and mechanonociception or thermosensation to mitigate deficits in touch (11). This is also true for human sexual touch (table S1). Nonetheless, the crucial role of PIEZO2 for perineal touch in mice and hu- mans may have therapeutic potential: Topical PIEZO2 inhibitors could provide targeted re- lief of genital hypersensitivity and pain, whereas agonists of PIEZO2 are candidates for alleviat- ing genital hyposensitivity. There are a number of limitations to this work. For example, PIEZO2 deficiency is extremely rare, and we were unable to carry out detailed quantitative sensory testing in a larger group of human subjects. Additionally, functional imaging experiments were carried out in anes- thetized mice, precluding evaluation of re- sponses during mating. Moreover, although we showed the necessity of gentle touch input for mating, we have not yet demonstrated the sufficiency of this sensory pathway for sexual function in awake behaving animals. We also anticipate that there are likely to be additional specialized roles for mechanosensory neurons in mating that were not revealed in this study. Even the very gentlest of perineal touches elicits a highly stereotyped startle reaction from mice that is easy to anthropomorphize (movie S1). This PIEZO2-dependent response is quite different from touch to other parts of the body, which typically evokes more modest reactions and does so only at much greater forces. PIEZO2- dependent perineal touch is also a crucial driver of successful mating both for male and female mice. Future studies should help define addi- tional subtypes of sensory neurons needed for sexually dimorphic reactions and how peri- genital sensation is organized in the spinal cord and brain to prioritize salience. Ultimately, how- ever, the profound impact of PIEZO2 deficiency that we describe provides a sensory basis at the molecular and cellular level for an aspect of life that throughout history has engaged in human imagination (42) and thought (1). RE FERENCES AND NOTES 1. C. Darwin, The Descent of Man and Selection in Relation to Sex (Murray, 1871), vols. 1 and 2. J. J. Dinsmore, Auk 87, 305–321 (1970). 2. 3. S. J. Gould, Evolution 28, 191–220 (1974). 4. M. A. Changizi, Q. Zhang, S. Shimojo, Biol. Lett. 2, 217–221 (2006). 5. B. D. Charlton, M. S. Martin-Wintle, M. A. Owen, H. Zhang, R. R. Swaisgood, R. Soc. Open Sci. 5, 181323 (2018). 6. V. S. Mandiyan, J. K. Coats, N. M. Shah, Nat. Neurosci. 8, 1660–1662 (2005). 7. L. Stowers, T. E. Holy, M. Meister, C. Dulac, G. Koentges, Science 295, 1493–1500 (2002). 8. C. C. Voigt et al., J. Mammal. 89, 1401–1410 (2008). 9. C. A. Kell, K. von Kriegstein, A. Rösler, A. Kleinschmidt, H. Laufs, J. Neurosci. 25, 5984–5987 (2005). 10. B. Coste et al., Science 330, 55–60 (2010). 11. A. T. Chesler et al., N. Engl. J. Med. 375, 1355–1364 (2016). I. Abdus-Saboor et al., Cell Rep. 28, 1623–1634.e4 (2019). 12. S. S. Ranade et al., Nature 516, 121–125 (2014). 13. Y. García-Mesa et al., J. Anat. 238, 446–454 (2021). 14. Y. García-Mesa et al., J. Anat. 239, 892–902 (2021). 15. 16. T. Hua et al., Nat. Neurosci. 23, 854–868 (2020). 17. S. E. Murthy et al., Sci. Transl. Med. 10, eaat9897 (2018). 18. V. E. Abraira, D. D. Ginty, Neuron 79, 618–639 (2013). 19. K. Nonomura et al., Nature 541, 176–181 (2017). 20. S.-H. Woo et al., Nat. Neurosci. 18, 1756–1762 (2015). 21. M. Szczot et al., Sci. Transl. Med. 10, eaat9892 (2018). 22. M. L. Sorrells et al., BJU Int. 99, 864–869 (2007). 23. C. B. Bleustein, H. Eckholdt, J. C. Arezzo, A. Melman, J. Urol. 169, 2266–2269 (2003). 24. L. Li et al., Cell 147, 1615–1627 (2011). 25. V. E. Abraira et al., Cell 168, 295–310.e19 (2017). 26. S. H. Woo et al., Nature 509, 622–626 (2014). 27. L. Madisen et al., Neuron 85, 942–958 (2015). 28. K. L. Marshall et al., Nature 588, 290–295 (2020). 29. F. A. Beach, Horm. Behav. 7, 105–138 (1976). 30. B. D. Sachs, Physiol. Behav. 24, 489–492 (1980). 31. N. Agarwal, S. Offermanns, R. Kuner, Genesis 38, 122–129 (2004). 32. N. Sharma et al., Nature 577, 392–398 (2020). 33. M. Q. Nguyen, C. E. Le Pichon, N. Ryba, eLife 8, e49679 (2019). 34. L. J. von Buchholtz et al., Neuron 109, 285–298.e5 (2021). 35. C. J. Burnett et al., eLife 8, e44527 (2019). 36. S. X. Zhang et al., Nature 597, 245–249 (2021). 37. X. Zhou et al., Science 379, 820–825 (2023). 38. R. D. Johnson, Z. Halata, J. Comp. Neurol. 312, 299–310 (1991). 39. D. Ohmori, Anat. Embryol. 70, 347–410 (1924). 40. D. Huzard et al., Sci. Adv. 8, eabo7566 (2022). 41. H. Olausson et al., Nat. Neurosci. 5, 900–904 (2002). 42. M. Vatsyayana, The Kama Sutra of Vatsyayana, R. F. Burton, B. Indrajit, S. P. Bhide, Transl. (1883). AC KNOWLED GME NTS We thank T. Delong, M. Singh, and M. Bradson for distributing questionnaires to human subjects and collating clinical notes; members of the clinical staff at NCCIH and Bönnemann groups for support in arranging human studies; and M. Szczot, N. Ghitani, D. Macdonald, and R. Hardy for input and expertise with animal experiments. Hormone analysis was carried out by the University of Virginia Center for Research in Reproduction Ligand Assay and Analysis Core. Funding: This work was supported by the National Institutes of Health, NCCIH Z01-ZIAAT000028 (to A.T.C.); National Institutes of Health, NIDCR Z01-ZIADE000561 (to N.J.P.R.); National Institutes of Health, NINDS Z01-ZIANS003129 (to C.G.B.); and the Howard Hughes Medical Institute (to A.P.) Author contributions: Conceptualization: R.M.L., N.J.P.R., and A.T.C. Methodology: R.M.L., M.F., L.J.v.B., E.F., C.G.B., N.J.P.R., and A.T.C. Investigation: R.M.L., M.F., J.O., E.F., M.R.S.-V., M.N., M.Q.N., M.J., and D.S. Funding acquisition: A.P., C.G.B., N.J.P.R., and A.T.C. Supervision: A.P., C.G.B., N.J.P.R., and A.T.C. Writing – original draft: R.M.L., L.J.v.B., N.J.P.R., and A.T.C. Writing – review and editing: R.M.L., M.F., L.J.v.B., E.F., A.P., C.G.B., N.J.P.R., and A.T.C. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data generated and/or analyzed during the current study are provided in the supplementary materials and/or have been deposited in Dryad. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal-article- reuse. This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the author- accepted manuscript (AAM) of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg0144 Materials and Methods Figs. S1 to S7 Table S1 References (43–46) Movies S1 and S2 Submitted 27 November 2022; accepted 13 July 2023 10.1126/science.adg0144 Lam et al., Science 381, 906–910 (2023) 25 August 2023 5 of 5
10.1126_science.adf5121
RES EARCH R E S E A R C H A R T I C L E S U M M A R Y ◥ MICROBIAL ECOLOGY Metabolic interaction models recapitulate leaf microbiota ecology Martin Schäfer†, Alan R. Pacheco†, Rahel Künzler, Miriam Bortfeld-Miller, Christopher M. Field, Evangelia Vayena, Vassily Hatzimanikatis, Julia A. Vorholt* INTRODUCTION: Microbiome composition is tied to host health and ecosystem function. However, the processes that determine spe- cies abundances in an environmental context remain poorly understood. Given that micro- biomes are bounded by resource availability, insights into the metabolic capacities of their constituent organisms present a key avenue for predicting interspecies interactions and broader microbiome assembly rules. Plants constitute the largest biomass on Earth and are colonized by a variety of micro- organisms that affect their health and growth. The composition of plant microbiomes has been found to be largely deterministic, sug- gesting the presence of defined drivers of community assembly. However, the metabolic mechanisms of the interactions underlying these patterns remain poorly understood. RATIONALE: To assess the degree to which pat- terns of resource allocation contribute to micro- bial interactions, we first profiled the carbon source–utilization capabilities of 224 represen- tative bacterial strains isolated from leaves of wild Arabidopsis thaliana plants. To do this, we cultivated each isolate on 45 different carbon sources and developed a computational tool to score the growth of each strain in an auto- mated way. This screen allowed us to calculate the degree of metabolic niche overlap between the strains, which informed predictions of re- source competition on the plant host. We further leveraged this experimental data to generate a collection of 224 genome-scale metabolic mod- els, which encompass the metabolic networks of each organism. We used these models to predict the outcomes of interspecies interactions on the plant host and to obtain mechanistic in- sight into specific patterns of resource use and exchange between these bacteria. RESULTS: We found that the carbon source– utilization profiles of the strains exhibited a strong phylogenetic signature, on the basis of both the experimental screen and the path- ways represented in the genome-scale models. We validated the performance of the models and of predictions based on niche overlap by conducting competition experiments with two sets of strains on the Arabidopsis host. These experiments revealed a dominance of negative interaction outcomes (i.e., a strain reached a lower overall population level upon cocoloni- zation compared with monoassociation), which was in agreement with predictions of high interstrain niche overlap. The genome-scale models provided an additional degree of insight into these interactions, also correctly predicting instances of positive outcomes observed in planta and further underscoring the importance of carbon metabolism to com- munity assembly. After this experimental validation, we modeled interaction outcomes for more than 17,500 pairs of strains. We predicted that in 94% of pairings, at least one strain would experience a reduction in abun- dance compared with monoassociation. We then analyzed the metabolic fluxes underlying these predicted outcomes, which revealed a high prevalence of competition for sugars. This competition could be offset by increased uptake of amino and organic acids in our simulations, which points to the metabolic mechanisms underpinning the positive inter- action outcomes that we observed in planta. CONCLUSION: Our results indicate a major role for carbon source availability and metabolic interactions in community assembly in the plant host, which, together with the trait con- servation observed here, likely contribute to the overall deterministic outcomes of plant microbiome assembly. In addition to recapi- tulating interspecies interactions in situ, our modeling framework represents a powerful tool for determining the metabolic mecha- nisms that lead to the emergence of specific ecological patterns. This knowledge will ulti- mately enable targeted microbiome design, which is pivotal to microbiome applications in health, agriculture, and the environment.▪ The list of author affiliations is available in the full article online. *Corresponding author. Email: jvorholt@ethz.ch †These authors contributed equally to this work. Cite this article as M. Schäfer et al., Science 381, eadf5121 (2023). DOI: 10.1126/science.adf5121 READ THE FULL ARTICLE AT https://doi.org/10.1126/science.adf5121 Carbon utilization profiles of leaf bacterial strains 45 carbon sources 224 strains Metabolic niche overlap and genome-scale models Predicted pairwise and community interaction outcomes Experimental validation in planta i A n a r t s f o e m o c t u o n o i t c a r e t n I Interaction types 23% 46% 2% 3% 1% 25% Growth No growth with strain or community B Negative Positive Competitive Parasitic Amensal Commensal Mutualistic Neutral Modeling of metabolic interactions predicts ecological outcomes for leaf microbiota members in situ. We mapped the resource-use capabilities of 224 representative bacterial strains of the A. thaliana leaf microbiome by testing their ability to grow on 45 different carbon sources. These data allowed us to predict interaction outcomes between these strains using genome-scale models and metrics of niche overlap. Experimental validation in planta underscored the importance of resource competition and suggested metabolic mechanisms underlying specific interaction outcomes. Schäfer et al., Science 381, 42 (2023) 7 July 2023 1 of 1 RES EARCH R E S E A R C H A R T I C L E ◥ MICROBIAL ECOLOGY Metabolic interaction models recapitulate leaf microbiota ecology Martin Schäfer1†, Alan R. Pacheco1†, Rahel Künzler1, Miriam Bortfeld-Miller1, Christopher M. Field1, Evangelia Vayena2, Vassily Hatzimanikatis2, Julia A. Vorholt1* Resource allocation affects the structure of microbiomes, including those associated with living hosts. Understanding the degree to which this dependency determines interspecies interactions may advance efforts to control host-microbiome relationships. We combined synthetic community experiments with computational models to predict interaction outcomes between plant-associated bacteria. We mapped the metabolic capabilities of 224 leaf isolates from Arabidopsis thaliana by assessing the growth of each strain on 45 environmentally relevant carbon sources in vitro. We used these data to build curated genome-scale metabolic models for all strains, which we combined to simulate >17,500 interactions. The models recapitulated outcomes observed in planta with >89% accuracy, highlighting the role of carbon utilization and the contributions of niche partitioning and cross-feeding in the assembly of leaf microbiomes. V arious hosts, including animals (1, 2), hu- mans (3, 4), and plants (5), support mi- crobial communities with hundreds to thousands of different species. Recent studies have revealed that a determi- nistic relationship exists between environmental composition and community structure, even for complex microbiomes (6–10). Thus, a thorough understanding of the resources available to a community could enable prediction of its spe- cies and functional composition. Not only would the ability to make such predictions help us bet- ter understand the relationship between envi- ronment and phenotype, but it would also provide an accessible way to rationally design synthetic communities with defined functions. Despite this potential, we still lack a com- prehensive understanding of the information necessary to predict whether a specific orga- nism will survive or be outcompeted in a parti- cular environmental context. Resolving this question remains an active area of research, with studies that use interaction outcomes between pairs of organisms to predict overall community behavior representing a particu- larly attractive approach (11–13). Nonetheless, conducting the necessary mapping of pos- sible interactions between all organisms re- mains experimentally challenging even for relatively simple communities and largely out of reach for complex host-associated micro- biomes in situ. Given these limitations, one may ask which ecological concepts can be used to predict the 1Institute of Microbiology, ETH Zurich, Zurich, Switzerland. 2Laboratory of Computational Systems Biotechnology, École Polytechnique Fédérale de Lausanne, EPFL, Lausanne, Switzerland. *Corresponding author. Email: jvorholt@ethz.ch †These authors contributed equally to this work. role of individual organisms within a commu- nity context. Microbial populations are fun- damentally bounded by resource availability, where limiting amounts of micronutrients such as carbon, nitrogen, and phosphorus act to constrain community size and lead to the emergence of competition between organisms (14–16). Although the prevalence of these com- petitive interactions has begun to be quanti- fied in a variety of ecosystems (12, 17–20), the general rules by which they influence community-wide assembly patterns remain unresolved. As such, a commonly used ap- proach is to infer the competitive potential between organisms from the degree of metab- olic niche overlap between them, which relies solely on knowing their individual resource- utilization capabilities (21, 22). Although a resource-by-resource understanding of the metabolic profiles of complex microbiomes can be challenging to obtain, it represents a more accessible approach for inferring inter- species interactions and has proven success- ful at predicting high-level patterns of species diversity in communities (10, 23–26). Despite its benefits, the predictive power of niche overlap is limited in that it cannot account for the emergence of positive ecolo- gical interactions between organisms either through resource partitioning, cross-feeding, orthogonal resource-acquisition strategies, or evolved cooperation (27–30). As a complementary approach, genome-scale metabolic models pres- ent a tractable way to integrate the effects of these additional mechanisms into predictions of community structure (31, 32). Genome-scale models are mathematical representations of the metabolic capabilities of individual organ- isms. They incorporate genes, reactions, and metabolites associated with a given organism’s metabolic network and are applied to quanti- tatively assess how organisms can use available resources for growth (32). In addition to gener- ating predictions of resource allocation for individual organisms, combinations of genome- scale models of different organisms have been increasingly used to mechanistically describe pairwise and community-wide dynamics (33–35). Combining genome-scale models with pre- dictions of niche overlap therefore presents a powerful way to predict interspecies interac- tions within microbiomes. In this study, we generated metrics of metab- olic niche overlap and a collection of genome- scale models to predict interaction outcomes between 224 bacterial members of the Arabidopsis thaliana leaf microbiome. This environment, known as the phyllosphere, is an oligotrophic habitat where competitive interactions between its resident microbes are prevalent (19, 36, 37). Because it contains a variety of different car- bon sources (38, 39), it is an ecosystem that is well suited for studying the effects of resource allocation on interspecies interactions and community structure. In particular, the exposed nature of the phyllosphere makes it a relatively accessible setting for study in controlled con- ditions. The microbiota of the phyllosphere has been shown to confer beneficial functions to the plant host (40–42), which, together with the scale of colonizable surfaces presented by leaves (37, 43, 44) and the economic impor- tance of crops (45, 46), underscores the value of studying the processes underlying its assem- bly (47, 48). Results Profiling the carbon source–utilization capabilities of phyllosphere bacteria Comprehensive collections of environmental strains are valuable resources for studying relevant host-microbe and microbe-microbe interactions (42, 49–52). To examine interaction outcomes among members of the Arabidopsis phyllosphere (Fig. 1), we used the At-LSPHERE: a collection of 224 strains isolated from leaves of wild A. thaliana plants that represent a cross section of the taxonomic and functional diver- sity of the phyllosphere microbiota (53). We first assessed the ability of each strain to grow on solidified minimal medium supplemented with individual carbon sources from a set of 45 (table S1). These carbon sources, comprising a range of sugars, organic acids, sugar alco- hols, one-carbon compounds, aromatic com- pounds, and amino acids, were selected on the basis of the high prevalence of resources such as glucose, sucrose, and some amino acids that are known to account for a substantial frac- tion of carbon available to leaf epiphytes (36, 38, 39, 54). These carbon sources were also selected in part on the basis of the known ability of certain leaf strains to metabolize aromatic compounds, cleave glycosidic bonds, Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 1 of 13 RES EARCH | R E S E A R C H A R T I C L E Data acquisition Model generation Prediction Validation Annotated genomes Genome-scale models Evaluate model performance l a n o i t a t u p m o C l a t n e m i r e p x E Carbon source utilization patterns Niche overlap negative interaction positive interaction Fig. 1. Overview of experimentally and computationally guided prediction and testing of metabolic interactions among bacterial members of the Arabidopsis phyllosphere microbiota. and/or utilize one-carbon compounds (37, 55). In addition to these carbon sources, we sup- plemented the media with vitamins to en- able the growth of auxotrophs (the set of carbon sources included amino acids, so we did not supply additional amino acids to the media). We used a plate assay to evaluate the growth of a given strain on a particular carbon source, which was compared with a carbon- free control, using both manual inspection and automated image processing (Materials and methods). Our in vitro metabolic screen revealed phy- logenetically contingent patterns of carbon- utilization capabilities across the At-LSPHERE collection (Fig. 2 and fig. S1). We found that strains belonging to the same genus displayed similar—or even identical—carbon source– utilization profiles, suggesting trait conserva- tion between closely related species (Fig. 2A). We also identified genus-specific signatures of resource utilization, as well as cases in which strains could not grow on any of the tested carbon sources (e.g., Chryseobacterium and Exiguobacterium). These latter strains most likely have nutritional requirements that are not met by the minimal medium we provided (e.g., amino acid auxotrophies), because they were able to produce visible colonies on a complex medium (R-2A+M, Materials and methods). By contrast, strains belonging to the genera Arthrobacter, Pseudomonas, and Rhizobium grew on a large number of carbon sources [27.5 ± 6.9, 27.1 ± 3.7, and 25.8 ± 3.6, respectively (mean ± SD)]. To quantify these differences, we assigned a metric of substrate versatility, V , to each strain, defined as the percentage of all 44 growth-yielding carbon sources that each organism was able to use for growth (for this analysis, methylamine was excluded because it did not support the growth of any strain). The most versatile strain was Arthrobacter sp. Leaf145 (V ¼ 77%), followed by Pseudomonas spp. Leaf15 and Leaf98, and Rhizobium sp. Leaf202 (V ¼ 73, 70, and 68%, respectively). When comparing the phylogenetic distribution of strain-specific substrate versatilities, we observed that Beta- and Gammaproteobacteria had above-average versatilities (Fig. 2B), whereas Actinobacte- ria (with the exception of Arthrobacter) and Bacteroidetes had below-average versatili- ties. Furthermore, our screen revealed the presence of a canonical niche occupied by methylotrophs that was distinct from and smaller than that of many other strains (Fig. 2A and fig. S2). Indeed, the lower average versatility of Alpha- and Betaproteobacteria— which include Methylobacterium and Methyl- ophilus spp., respectively—relative to Gam- maproteobacteria reflects the low versatility of methylotrophs. Our screen also revealed the varying degrees to which each resource promotes growth (Fig. 2C). To quantify these patterns, we assigned a measure of substrate fertility, F, to each tested carbon source, defined as the percentage of strains that was able to grow on that particular carbon source. The most fertile or commonly used carbon source was glucose (F ¼ 81% of strains), followed by succinic acid (F ¼ 80%) and glutamate (F ¼ 78%) (Fig. 2C and table S2). Among the most rarely used carbon sources were aromatic compounds, including trypto- phan (F ¼ 9%) and coniferyl alcohol (F ¼ 8%), a common component of lignin. Given the largely orthogonal metabolic niche occupied by methylotrophs relative to all other strains, we repeated our calculations of substrate fertility without methylotrophs to more clearly detect patterns of fertility among the other substrates. This analysis revealed that mono- and disaccharides build a core set of carbon sources commonly used by the large majority (F ≥ 75%) of tested strains and that the most fertile substrate, glucose, was con- sumed by 95% of the strains. The next most common group of carbon sources (≥ 50%) included organic acids and amino acids that feed into the tricarboxylic acid (TCA) cycle (fig. S3). These experimentally determined carbon source–utilization capabilities allowed us to calculate the degree of niche overlap among our strains. We defined a niche overlap index (NOI) for each focal strain, A, with respect to any other strain, B, as the set of usable carbon sources shared by the focal strain and the competitor divided by the total number of car- bon sources used by the focal strain (23, 56): NOIA;B ¼ NC Sources; A ∩ B NC Sources; A We calculated NOI values for all pairs of strains, setting a threshold of 75% as likely prone to competitive exclusion (57) (Fig. 2D). In accordance with the similar carbon source– utilization profiles, we found high niche over- lap within most genera, suggesting frequent intragenus competitive interactions. Addition- ally, we found that methylotrophic strains form two interconnected interaction spaces: one comprising Methylobacterium and the other comprising Methylophilus strains, underscoring Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 2 of 13 RES EARCH | R E S E A R C H A R T I C L E A Monosaccharide Disaccharide Sugar alcohol Organic acid C1 compound Alcohol Aromatic Amino acid Alphaproteobacteria Betaproteobacteria Gammaproteobacteria Bacteroidetes Deinococcus−Thermus Actinobacteria Firmicutes At-LSPHERE strain paired with strain B D i A n a r t s f o I O N 25 50 Substrate versatility [%] 75 100 B Alphaproteobacteria Betaproteobacteria Gammaproteobacteria Actinobacteria Bacteroidetes Firmicutes Deinococcus−Thermus 0 C Monosaccharide Disaccharide Organic acid Amino acid C1 compound Alcohol Aromatic compound 0 25 50 Substrate fertility [%] 75 100 Rhizobium Arthrobacter Methylobacterium Methylophilus Microbacteriaceae NOI 0 0.25 0.5 0.75 1 Fig. 2. Carbon source–utilization screen reveals hotspots of high nutritional niche overlap. (A) Carbon source–utilization map of 224 At- LSPHERE strains. Fill color of the boxes indicates growth (black) or no growth (gray) for each strain on the x axis on a given carbon source on the y axis. Top-level clustering for strains reflects phylogeny based on full-length 16S ribosomal RNA gene sequences. Colored bars indicate phylum or Proteobacterium class and correspond to the order in the legend. Carbon sources are sorted by compound groups. (B and C) Density plots showing (B) strain-specific substrate versatility grouped by phylum or Proteobacterium class (corresponding to the order in the legend) and (C) substrate fertility grouped by compound group. Black dots indicate individual values, and the red vertical line shows the median for each group. Substrates in classes C1 and aromatic compounds that also fall into another substrate group are only shown once and omitted from the other group. (D) Heatmap of niche overlap indices (NOI) for all binary combinations of 215 At-LSPHERE strains. The color of the tile indicates high (>0.75, blue shades) or low (<0.75, gray shades) degree of niche overlap for the strain on the y axis in combination with the strain on the x axis. Strains are sorted by phylogeny, and the colored bars indicate the phylum or Proteobacterium class. NOI was only calculated for strains that grew on ≥1 carbon source in vitro. Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 3 of 13 RES EARCH | R E S E A R C H A R T I C L E their degree of metabolic specialization com- pared with all other strains. Moreover, we found that members of the Microbacteriaceae had high niche overlap with a high number of other taxa, largely owing to their low substrate versatilities. We also observed the opposite case: Rhizobium strains had a low NOI with other taxa, whereas essentially all strains (ex- cept Methylobacteria) had a high NOI with Rhizobium strains. An in silico representation of phyllosphere bacterial communities Having screened the carbon source–utilization capabilities of the At-LSPHERE collection, we sought to model how these bacteria would interact metabolically in a leaf environment. To do this, we adopted a stoichiometric metab- olic modeling approach because it allowed us to explicitly simulate strain- and substrate- specific patterns of resource utilization, con- version, and exchange in a multitude of defined environmental conditions. We thus began by creating draft genome-scale metab- olic reconstructions (58) for each member of the At-LSPHERE. This process yielded a separate metabolic network for each of the 224 At-LSPHERE strains, containing all the metabolic reactions predicted to be contained by each organism on the basis of its genome annotation. After generating these draft metab- olic reconstructions, we used flux balance analysis (31) to simulate the growth of each strain individually within a complete medium containing all 45 carbon sources used in our in vitro screen. We found that whereas some of our reconstructions were predicted to pro- duce biomass when using this complete me- dium, the majority could not do so when only individual carbon sources were provided (Fig. 3A). Although this inability to grow was at odds with our experimental results, it was not unexpected because models generated solely from genomic information are often missing key metabolic pathways and thus have limited predictive power without additional experi- mental curation (33, 59–61). We therefore used the results of our carbon source screen to fill the apparent gaps in each metabolic reconstruc- tion by using the tool NICEgame (62), which suggests biologically relevant alternative path- ways to account for missing reactions by lever- aging information on reaction thermodynamics. We followed this step with additional manual curation (Materials and methods, table S3, and fig. S4) and then compared each of the resulting 224 models with their correspond- ing strains on the basis of their ability (or in- ability) to use the 45 carbon sources selected for our in vitro screen (fig. S5). This process resulted in each genome-scale model having a final balanced accuracy of 0.98 ± 0.07 (Fig. 3A) (mean precision 1, mean recall 0.96), with 214 models (96%) having a balanced accuracy of at least 0.85, and 189 models (84%) being 100% accurate. Our curation process yielded a collection of 224 genome-scale models that revealed further metabolic capabilities and physiologi- cal characteristics of the At-LSPHERE collec- tion (Fig. 3). The models contained between 1568 (Flavobacterium sp. Leaf359) and 3004 (Pseudomonas sp. Leaf15) reactions (Fig. 3B), which moderately correlated with the genome size of their corresponding strain [coefficient of determination (R2) = 0.54, fig. S6A]. Reac- tions involved in the biosynthesis of glycero- phospholipids, key components of the cell membranes of Gram-negative bacteria, were enriched in members of the Proteobacteria within our collection (Fig. 3C). We also found Firmicutes and Actinobacteria to contain more reactions related to degradation of plant polysaccharides (e.g., melibiose, raffinose, and sucrose) relative to other taxonomic classes and identified an enrichment for benzoate degradation capabilities as previously reported in Betaproteobacteria (63, 64). The high degree of similarity to our strains’ known carbon source–utilization capabilities was supported by a clustering analysis of the reactions con- tained in each model (Fig. 3D). In this analysis, Methylobacteriaceae spp. and Methylophilaceae spp. formed clusters distinct from most other models, and Acidovorax strains clustered to- gether with the majority of Pseudomonas and Rhizobium organisms. These clusters largely mirrored those generated from our in vitro data alone (Fig. 3E), which underscored the phylogenetic contingency of resource-utilization patterns and may underlie the deterministic assembly of microbiomes in general. With this in silico representation of the At-LSPHERE col- lection, we proceeded to use our genome-scale models to simulate the ecological outcomes of mixed-culture experiments. Computational models recapitulate ecological patterns in planta We assessed the predictive power of the genome-scale models and the NOI against experimental data in two in planta experi- ments (Fig. 4, A and B). In a first experiment, we inoculated A. thaliana seedlings with pair- wise combinations of seven bacterial strains, or with a community consisting of all seven strains (referred to as SynCom7), and com- pared the colonization level of each strain in combination with the colonization density achieved in monoassociation (Fig. 4B). To represent a range of metabolic capabilities, we selected strains across a gradient of sub- strate versatility values and hence a gradient of NOI. For the first experiment, we selected representative strains from the highly versatile Arthrobacter spp., Pseudomonas spp., and Rhizobium spp. (Leaf145, V ¼ 77%; Leaf15, V ¼ 73%; and Leaf202, V ¼ 68%; respec- tively), along with four additional strains with intermediate-to-low versatility values: Rhodo- coccus sp. Leaf233, Sphingomonas spp. Leaf34 and Leaf257, and Microbacterium sp. Leaf179 (V ¼ 55, 36, 45, and 41%, respectively). On the basis of NOI, we expected the strains with lower versatilities to more frequently de- crease in abundance upon cocolonization with other strains, either in pairwise combinations or in a community context (Fig. 4C). How- ever, all other strain pairings had low NOI values (<75%), limiting the degree to which we could infer interaction outcomes for these combinations. We used the genome-scale models corres- ponding to these seven strains to simulate their growth within an environment reflecting the carbon source availability of the Arabidopsis phyllosphere (Materials and methods). To obtain an indicator of the ecological outcomes of these pairings, we compared the predicted growth of each strain on its own with that in coassociation with another: A strain was deemed to have experienced a negative inter- action outcome if its biomass production rate in coassociation was lower than that on its own, and a positive outcome if its biomass production rate was higher in coassociation. These simulations predicted an approximately even distribution of positive and negative outcomes for all seven organisms (Fig. 4C). However, when we considered the magnitude of the changes in growth experienced by each strain, we found that predictions skewed more strongly negative than positive. This observa- tion is consistent with previous descriptions of the prevalence of competitive interactions in the phyllosphere (18, 19), which suggests the commonality of exclusion of one strain by another. Indeed, when we compared the growth of each strain on its own with that in the presence of all six other strains, we predicted almost exclusively negative outcomes. Despite this overall distribution, we found that the number of positive or negative outcomes dif- fered depending on the strain in question, with Microbacterium sp. Leaf179 and Sphingomonas sp. Leaf257 almost always experiencing a re- duction in growth when paired with another strain. By contrast, Rhizobium sp. Leaf202, Pseudomonas sp. Leaf15, and Arthrobacter sp. Leaf145 were most often predicted to benefit from coinoculation with another of our selected strains. These predicted positive outcomes corresponded to instances of low NOI values (Fig. 4C), suggesting that these strains could metabolically evade competition by a partner strain and may additionally bene- fit from changes in the ecosystem (e.g., cross- feeding) induced by the partner strain. The interaction outcomes of these seven strains in planta were largely congruent with those predicted by NOI and metabolic model- ing. Of the outcomes that showed a fold change Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 4 of 13 B 1 60 0 M e a n n u m b e r p e r m o d e l i n o f t a x o n r e a c t i o n s RES EARCH | R E S E A R C H A R T I C L E A y c a r u c c a d e c n a a B l Initial Curated 0.4 0.6 0.8 C Plant polysaccharide degradation Oxidative phosphorylation Methane metabolism Selenoamino acid metabolism Biotin metabolism tRNA charging Nitrogen metabolism Beta-alanine metabolism Ascorbate and aldarate metabolism Benzoate degradation Glutathione metabolism Sulfur metabolism Nucleotide salvage pathway ROS detoxification Cholesterol metabolism Inositol phosphate metabolism Peptide metabolism Vitamin B12 metabolism Galactose metabolism Propanoate metabolism Terpenoid backbone biosynthesis Heme synthesis Urea cycle Aminosugar metabolism Thiamine metabolism Tyrosine metabolism Ubiquinone and other terpenoid-quinone biosynthesis Alanine and aspartate metabolism Fructose and mannose metabolism CoA synthesis Pentose and glucuronate interconversions Vitamin B6 metabolism Starch and sucrose metabolism Purine catabolism Histidine metabolism Purine synthesis Phenylalanine metabolism Butanoate metabolism Energy metabolism Glyoxylate and dicarboxylate metabolism Tryptophan metabolism Cell wall biosynthesis Glutamate metabolism Lysine metabolism Vitamin B2 metabolism NAD metabolism Pyrimidine synthesis Citric acid cycle Fatty acid oxidation Arginine and proline metabolism Pyruvate metabolism Pentose phosphate pathway Methionine and cysteine metabolism Folate metabolism Pyrimidine catabolism Glycine, serine, alanine, and threonine metabolism Fatty acid synthesis Glycolysis/gluconeogenesis Valine, leucine, and isoleucine metabolism Glycerophospholipid metabolism 2 2 63 83 83 4 02 01 7 3 5 73 73 4 2 3 0 2 4 2 2 0 82 1 9 8 2 0 5 3 6 3 1 6 8 2 8 0 0 2 4 4 6 4 2 3 3 4 4 4 5 4 0 6 4 7 2 4 6 8 3 3 5 4 1 9 3 1 7 3 1 4 3 3 8 3 4 8 3 6 0 3 1 2 3 1 2 1 6 3 52 5 33392314122125733321629 91123 524226762 10407 42 4 25 26 2 76 191 84 78 73 160 220 274 265 267 400 126 61 139 177 416 459 414 408 98 15 59 434 48 129 127 58 83 130 51 50 53 131 148 70 313 404 394 405 201 180 82 359 170 132 250 41 194 176 216 189 1084663999212211912190123456118361344401396202167681 86 91 93 10685 99 469 112 87 102 100 113 117 104 125 89 88 465 94 111 326 72 182 187 196 49 75 13 406 380 369 354 278 233 7 247 258 225 446 285 307 374 350 289 245 272 291 3 395 334 234 145 137 141 69 337 261 6 3 4 9 7 1 1 5 1 3 0 2 8 8 2 7 4 3 1 5 3 0 2 3 9 5 1 1 6 1 11 7 1 4 1 3 5 2 3 4 6 2 2 2 2 6 3 3 8 4 2 92 9 9 6 2 2 9 4 1 8 5 2 1 0 3 3 5 2 4 4 2 8 3 4 1 5 2 5 4 1 8 3 1 5 4 3 0 4 1 6 4 1 7 2 2 6 3 81 8 6 4 4 Alphaproteobacteria Betaproteobacteria Gammaproteobacteria Bacteroidetes Actinobacteria Firmicutes Deinococcus-Thermus No. of reactions 1500, 2500 Alphaproteobacteria Gammaproteobacteria Actinobacteria Betaproteobacteria Firmicutes Bacteroidetes Deinococcus-Thermus 0.5 1 2.9 7.9 0 0.73 Balanced accuracy Genome size (Mb) Model versatility D 1200 800 400 0 ) . r a v . l p x e % 9 . 3 1 ( 2 o C P -400 E 20 10 0 -10 ) . r a v . l p x e % 5 . 5 3 ( 2 o C P Acidovorax Methylobacterium Methylophilus Xanthomonas Sphingomonas Pseudomonas Rhizobium Microbacterium All others -1000 -500 0 500 1000 -20 -10 0 10 20 PCo1 (22.8% expl. var.) PCo1 (46.3% expl. var.) Fig. 3. Overview of collection of metabolic models for 224 bacterial members of the A. thaliana leaf microbiota. (A) Distributions of balanced accuracy as tested on 45 carbon sources for models generated solely from genomic information (top) and after curation (bottom). (B) Attributes of models as clustered by taxonomic phylum and class. Numbers bordering phylogenetic tree denote strain identity (e.g., 220 corresponds to Xylophilus sp. Leaf220). The innermost ring represents the balanced accuracy of each model, the second ring indicates the size of the genome used to generate each model (53), and the third ring represents the versatility, V, of each model. The outermost bars represent the number of reactions contained in each model (minimum = 1568, maximum = 3004). (C) Clustered heatmap of 60 most highly represented reaction types, scaled by the number of models corresponding to each phylum and class. (D and E) Principal coordinates analysis (PCo) of strains as determined by (D) reactions in genome-scale models and (E) in vitro carbon source screen, with select genera highlighted. greater than 2 or less than 0.5, we found that a majority were negative, confirming the dom- inance of competition between these strains (Fig. 4D). In the SynCom7 condition, a sig- nificant abundance reduction was observed for all strains except for Rhizobium sp. Leaf202, matching our predictions and supporting the presence of increased competitive pressure in a community context. Our experimental data also highlighted the predictive power of NOI data alone for predicting interaction outcome directionality, with 93% of instances of nega- tive outcomes corresponding to an NOI >0.75. Predictions generated by the genome-scale models yielded an increased degree of gra- nularity, recapitulating the significant inter- action directionalities that we observed in planta with a balanced accuracy of 89% (tables S4 and S5). These predictions additionally captured the weaker directionalities of border cases that did not meet our experimental Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 5 of 13 RES EARCH | R E S E A R C H A R T I C L E A B n o i t i a c o s s a o n o m n o x u l f s s a m o B i s t n u o c U F C Δt i t i a n b m o c n i mono association in combination mono association in combination C Leaf34 Leaf179 Leaf257 Leaf233 Leaf202 Leaf15 Leaf145 E Leaf8 Leaf154 Leaf164 Leaf304 Leaf202 Leaf145 4 3 f a e L 9 7 1 f a e L 7 5 2 f a e L 3 3 2 f a e L 2 0 2 f a e L 5 1 f a e L 5 4 1 f a e L 7 m o C n y S NOI GSM NOI GSM 0 0.25 0.5 0.75 1 −5.0 −2.5 0.0 2.5 5.0 D Leaf34 Leaf179 Leaf257 Leaf233 Leaf202 Leaf15 Leaf145 F Leaf8 Leaf154 Leaf164 Leaf304 Leaf202 Leaf145 4 3 f a e L 9 7 1 f a e L 7 5 2 f a e L 3 3 2 f a e L 2 0 2 f a e L 5 1 f a e L 5 4 1 f a e L 7 m o C n y S |log2FC| < 1 padj ≤ 0.05 log2 FC −4 −2 0 2 4 f 8 a e L 4 5 1 a e L f 4 6 1 f a e L 4 0 3 a e L f 2 0 2 f a e L 5 4 1 f a e L 3 m o C n y S f 8 a e L 4 5 1 a e L f 4 6 1 f a e L 4 0 3 a e L f 2 0 2 f a e L 5 4 1 f a e L 3 m o C n y S Fig. 4. Model predictions and in planta validation of interactions. (A and B) Schematic overview of (A) interaction simulations using genome-scale models and (B) phyllosphere inoculation and interaction mapping procedure. Interactions are inferred by comparing a strain’s growth in combination with another strain with that in monoculture. This is computed as the log2 ratio of biomass fluxes for the genome-scale models or as the log2-fold change in CFUs per gram plant fresh weight for the in planta experiments. (C and E) Predicted interaction outcome for the strain indicated on the y axis in combination with the strain or SynCom indicated on the x axis, based on niche overlap index (NOI, top left) and genome scale models (GSM, bottom right). The fill color indicates high (>0.75, blue shades) or low (<0.75, gray shades) degree of niche overlap and predicted positive (red shades) or negative (blue shades) interactions based on genome-scale metabolic models. (D and F) Interaction outcomes observed in planta. Heatmap showing the log2-fold changes (log2FCs) (pairwise or SynCom inoculation versus monoassociation) for the strain on the y axis in combination with the strain or SynCom indicated on the x axis, based on absolute abundances obtained by CFU enumeration (n ¼ 11 to 12). The color of the boxes reflects the observed log2FC, and the black frames around the boxes indicate a significant difference compared with the monoassociation condition (two-sided Wilcoxon rank sum test, Holm-adjusted P ≤ 0.05). |Log2FC| < 1 are overlaid with a crosshatched pattern. Confusion matrices comparing modeling predictions and in planta outcomes are provided in table S4, and colony counts for the in planta experiments are provided in table S5. cutoff, as seen in the positive interaction out- come for Sphingomonas sp. Leaf34 paired with Microbacterium sp. Leaf179, of Arthrobacter sp. Leaf145 paired with Rhizobium sp. Leaf202, and of Rhizobium sp. Leaf202 in the SynCom7 condition. These weakly positive effects occurred despite high predicted NOI, suggesting the in- fluence of additional metabolic mechanisms not encompassed by competition. We sought to further validate our predic- tions with a second experiment spanning more extreme degrees of niche overlap. We generated pairings including the two highly versatile strains Rhizobium sp. Leaf202 and Arthrobacter sp. Leaf145 (V ¼ 68% and 77%, respectively) as well as the low-versatility strains Frigoribacterium sp. Leaf8, Curtobac- terium sp. Leaf154, Rathayibacter sp. Leaf164, and Frondihabitans sp. Leaf304 (V ¼ 27% for all four strains). All pairings resulted in a high degree of niche overlap for the low-versatility strains, whereas the NOI for either Leaf202 or Leaf145 with any other strain was low. Cor- respondingly, the genome-scale models pre- dicted almost exclusively negative interaction outcomes for all low-versatility strains, with weakly positive outcomes for Leaf202 in com- bination with all strains except Leaf145, as well as stronger positive outcomes for Leaf145 in combination with most other strains (Fig. 4E). We additionally predicted interaction out- comes for each of the three low-versatility strains paired with a combination of Leaf202 and Leaf145 (referred to as SynCom3), which resulted in strongly negative interaction out- comes for each strain. These predictions were tested with in planta experiments for these pairings: As predicted, all significant instances of log2-fold changes greater than 1 were nega- tive, including those for each low-versatility strain in the SynCom3 condition (Fig. 4F). Additionally, Leaf202 and Leaf145 each exper- ienced two instances of weakly positive out- comes when paired with other low-versatility strains, which were captured by our genome- scale modeling predictions. Our experiments confirmed the validity of the computational predictions (tables S4 and S5) and highlighted the strong contribution of resource competition to strain-specific interaction outcomes in situ. To test the specificity of our predictions to a leaf environment, we carried out a series of in vitro cultivations in which we cultured the same strain pairs and communities in shake flasks (Materials and methods). These culti- vations resulted in exclusively negative inter- action outcomes for all strains in the SynCom7 and SynCom3 conditions (fig. S7), similarly to our in planta experiments. These outcomes, which are consistent with an additional set of community simulations that used randomly selected strains (fig. S8), highlight the increased degrees of competition that organisms may be subject to in multispecies settings. However, Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 6 of 13 RES EARCH | R E S E A R C H A R T I C L E A Leaf145 + Leaf8 Leaf145 + Leaf233 C -4 -2 0 -4 -2 0 log2FC biomass flux Leaf145 Leaf8 Leaf233 B Sugars Amino acids Organic acids -4 -2 0 -4 -2 0 log2FC uptake flux (mmol/gDW/hr) Leaf145 Leaf8 Leaf233 Strongly negative Strongly positive E 23% 23% 2% 1% 3% 2 25% D y c n e u q e r f e m o c t u O 4000 2000 0 Outcome Positive Neutral Negative No growth i A n a r t s f o e m o c t u o n o i t c a r e t n I -8 -4 0 4 log2FC biomass flux 46% Competitive (-,-) Parasitic (-,+) Amensal (-,0) Commensal (+,0) Mutualistic (+,+) Neutral (0,0) in combination with strain B Alphaproteobacteria Betaproteobacteria Gammaproteobacteria Bacteroidetes Deinococcus-Thermus Actinobacteria Firmicutes F 6000 88.6% 50.8% 7.6% 37.0% y c n e u q e r F 0 0 Negative outcome Positive outcome 6000 90.7% 27.5% 8.5% 70.1% Negative outcome Positive outcome 6000 78.4% 38.3% 14.3% 49.9% Negative outcome Positive outcome 2 4 6 Ratio of total sugar uptake flux 0 0 8 2 4 6 Ratio of total amino acid uptake flux 0 0 8 2 4 6 8 Ratio of total organic acid uptake flux Fig. 5. Model-predicted interaction outcomes and mechanisms. (A and B) Log2FCs of (A) biomass and (B) resource uptake fluxes for two representative interactions validated in planta. Dots indicate absolute log2FCs of less than 0.05 mmol gDW–1 hour–1. (C) Predicted pairwise interaction outcomes between all 188 nonmethylotrophic strains in the At-LSPHERE (n ¼ 35,156 outcomes for 17,578 pairs). Hierarchical clustering was performed on interaction outcomes, with strain-specific phylogeny highlighted. White cells denote instances of no predicted growth in both mono- and coculture. (D) Distribution of pairwise interaction outcomes (n ¼ 35,156). Dashed lines separate outcomes in which a strain’s predicted biomass flux in coculture was either less than 90% of that in monoculture (strongly negative), within 10% of that in monoculture (neutral), or more than 110% of that in monoculture (strongly positive). (E) Classification of pairwise ecological outcomes (n ¼ 17,578). (F) Distributions of flux ratios between resource uptake in coculture and monoculture, according to corresponding interaction outcome. Only simulations in which a strain achieved growth in both monoculture and coculture are considered (n ¼ 28,316 outcomes). Differences between uptake rates of resource types provided in the simulated medium are highlighted for sugars (left), amino acids (center), and organic acids (right). Distributions of uptake fluxes are statistically significant for all three resource types (p << 1 (cid:2) 10 horizontal axes are truncated and show 98.8% of outcomes for sugars, 98.7% for amino acids, and 94.7% for organic acids. The dashed line at ratio of 1 separates instances of lower or higher uptake flux between coculture and monoculture, with percentages highlighting the number of instances less than or greater than 1. (cid:3)10) as determined by one-tailed Mann-Whitney U tests. For clarity, our in vitro experiments also resulted in sub- stantially fewer positive outcomes of pairwise interactions when compared with our in planta results, suggesting competitive pressures in- herent to batch cultures that are not captured by our use of NOI and genome-scale models. In particular, the lack of spatial structure may favor strains that experience fast growth when substrate availability is high and thus out- compete slower-growing strains within the timescale of a batch experiment. Moreover, this rapid depletion of nutrients contrasts with the resource dynamics of leaf surfaces, which exhibit a steady resupply of resources that can be accessed by epiphytic microbes (38). Al- though we do not explicitly consider spatial structure in our use of NOI and genome-scale models, these tools generate predictions based on a broader consideration of the various re- sources that can be used by the organisms at steady state. This assumption may therefore better reflect a broader and more continuously supplied pool of resources, which can be used by microbes on a population level in spatially structured settings. Compensatory metabolic mechanisms offset resource competition In addition to predicting interaction outcome directionalities, we used genome-scale model- ing to explore the metabolic mechanisms that could be underlying the observed ecological patterns. We first examined experimentally validated interactions to determine changes in resource uptake rates that emerged as a result of pairing two strains together. As rep- resentative examples, we looked specifically at two interactions involving Arthrobacter sp. Leaf145, a highly versatile strain that experi- enced a weakly positive effect when paired with Frigoribacterium sp. Leaf8 and a negative effect when paired with Rhodococcus sp. Leaf233 (Fig. 4, C to F, and Fig. 5A). Our flux balance simulations had predicted that in both cocul- tures, Leaf145 would have a lower net uptake flux of sugars compared with those experi- enced in monoculture (Fig. 5B). This was also the case for Leaf8, which was additionally pre- dicted to experience a reduction in amino acid uptake flux when paired with Leaf145. Whereas a similar reduction in amino and organic acid uptake occurred for Leaf233, Leaf145 was able to take up amino acids at rates similar to those noted in monoculture. The contribution of this reallocation of resources to the dominance of Leaf145 can be seen in its interaction with Leaf8, in which Leaf145 was able to shift its Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 7 of 13 RES EARCH | R E S E A R C H A R T I C L E metabolism to take up a slightly greater quan- tity of amino acids, as occurred in monoculture. The increased availability of these resources, in part a product of the low metabolic activity of Leaf8, suggests metabolic cross-feeding as a contributing factor in the positive effect ex- perienced by Leaf145 in coculture. The emergence of distinct resource-allocation patterns between these strain pairs prompted us to ask how widespread they could be across a wider ecosystem. We thus used the genome- scale models to generate predictions of pairwise interaction outcomes for all nonmethylotrophic strains in the At-LSPHERE, carrying out a total of 17,578 pairwise simulations (Fig. 5C). These simulations revealed the high prevalence of competitive effects, with 63.2% of outcomes predicted to be negative (Fig. 5D). Of these, the large majority (representing 58.2% of all outcomes) were predicted to be strongly nega- tive, which we defined as a strain’s biomass production rate in coculture being less than 90% of that in monoculture. Conversely, 25.5% of all outcomes were predicted to be strongly positive (biomass production in coculture more than 110% of that in monoculture), further un- derscoring the strength of competitive pres- sures within leaf bacterial communities. We combined these strain-specific effects to de- fine interaction outcomes between pairs, which revealed that in 94% of interactions, at least one participant experienced a negative out- come (Fig. 5E). As such, cooperative inter- actions were predicted to be relatively rare, with 3% of all pairwise interactions resulting in commensalism and only 1% resulting in mutualism. An analysis of metabolic fluxes across our simulations revealed key differences in rates of resource uptake between positive and nega- tive interaction outcomes (Fig. 5F). As expected, the models predicted significantly lower re- source uptake rates for organisms that de- creased in abundance relative to monoculture than for organisms that increased in abun- dance. However, whereas nearly all flux ratios associated with negative outcomes were below 1 (occurring for sugars, amino acids, and or- ganic acids in 88.6, 90.7, and 78.4% of nega- tive outcomes, respectively), many positive interaction outcomes displayed increases in amino acid and organic acid uptake flux (in 70.1 and 49.9% of positive outcomes, respec- tively). These patterns thus suggest a stron- ger degree of competition for sugars among phyllosphere bacteria, which is more likely to be offset by increased uptake of amino and organic acids. These predicted compensatory mechanisms suggested that strains with higher substrate versatilities would be more likely to increase in abundance when in the presence of another strain. Indeed, strains with very high versatil- ities (e.g., Acidovorax spp. and Pseudomonas spp.) were predicted to have higher than aver- age rates of experiencing positive outcomes (fig. S9). However, strain-specific substrate versatility itself was a poor predictor of posi- tive outcome frequency ( R2 ¼ 0.10) because strains with very low versatilities (e.g., Chrys- eobacterium spp., which did not grow on any of the supplied carbon sources individually, or Bosea sp. Leaf344) were also predicted to ex- perience strongly positive outcomes in rare cases. These effects suggest a role of cross-fed amino or organic acids as valuable sources of metabolic complementation (65), which in ad- dition to enabling highly versatile strains to outcompete a partner strain, can also facil- itate the survival of specialized organisms. Discussion Understanding the factors that contribute to the assembly of microbiomes remains a chal- lenge for the study of natural ecosystems. In this study, we have shown how predictions based on the resource-utilization capabilities of individual bacteria enable recapitulation of interaction outcomes observed in the phyllo- sphere. These predictions underscored the prevalence of competitive interactions between leaf strains. Additionally, the high accuracy (Fig. 4) of these predictions suggested that competition for resources is responsible for many outcomes observed in situ, as also re- cently observed in community contexts (19). In particular, our experiments showed how high degrees of niche overlap were reliably predictive of negative interaction outcomes in planta. This relationship demonstrates the role that highly versatile organisms play in communities, as well as the implications for the design of microbial consortia based on co- operative interactions, where pairings of highly competitive organisms can be avoided with greater certainty. Our work highlights how a resource-by-resource understanding of an orga- nism’s catabolic potential in monoculture is sufficient for making high-level predictions of competition in an ecologically relevant setting. In addition to mapping the metabolic ca- pabilities of a large strain collection at high resolution, our in vitro carbon source screen informed the curation of genome-scale meta- bolic models for each of the 224 strains. Our experimentally curated collection of genome- scale models allowed us to go beyond the po- tential of niche overlap by predicting positive interaction outcomes and suggesting their molecular mechanisms. Our curation process provided a quantitative basis for the need to incorporate experimental data into model generation, given the low accuracy of the ini- tial draft models, which were based on ge- nome annotations alone. Although improved annotation resources, as well as organism- specific biomass compositions, may serve to reduce this source of uncertainty (60, 61), we expect experimental curation to remain es- sential for selecting the optimal combination of gap-filled reactions to recapitulate an orga- nism’s exact resource-utilization profile. Our metabolic modeling predicted interac- tion outcomes on the basis of carbon source– utilization capabilities. Further curation that integrates additional properties such as vita- min auxotrophies and storage (66), as well as metabolic shifts occurring from gene regula- tion (67, 68), is likely to improve the quantitative predictive power of this framework (33, 69). Moreover, although our predictions based on metabolic mechanisms are informative of key aspects underlying the assembly of plant- associated microbiomes, they do not consider additional factors that are known to shape leaf communities, such as signaling mole- cules, antagonistic interactions, or host immu- nity (5, 70–73). A potential consequence of this limitation may be seen in the example of Pseudomonas sp. Leaf15, which experienced a strong negative interaction in planta when paired with Rhizobium sp. Leaf202 (Fig. 4, C and D). Even though this outcome was qual- itatively predicted by our simulations, the pres- ence of a type VI secretion system in Leaf202, previously described to be active against a Pseudomonas strain (42), may also be produc- ing an additional inhibitory effect not captured in our modeling framework. Furthermore, al- though our experimental approach that quan- tified community composition on the basis of bulk sampling aligns with our use of flux balance analysis, it abstracts away the influ- ence of spatial structure on interaction out- comes. Leaf surfaces are known to exhibit considerable nutrient heterogeneity, which, together with the aggregation of microbes around stomata or in microdroplets, is likely to affect interaction directions at the micro- scale (36, 74, 75). Capturing varying degrees of uncertainty in our modeling predictions can partially represent this heterogeneity (fig. S10), but an improved understanding of the nutrient compositions at specific locations on the leaf may enable future parametrization of existing metabolic modeling tools that can explicitly consider spatial structure (76–79). Our present results underscore the strengths of an integrated approach for generating eco- logical predictions in a mechanistic yet scal- able way, while establishing a computational resource for further exploration of the molec- ular mechanisms that underlie the assembly of complex microbiomes. Materials and methods Cultivation of bacteria At-LSPHERE strains were grown on R-2A agar (Sigma-Aldrich) supplemented with 0.5% (v/v) methanol (R-2A+M) at 22°C. Strains were rou- tinely streaked out from a cryo stock stored at −80°C, grown for 4 days, streaked on fresh Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 8 of 13 RES EARCH | R E S E A R C H A R T I C L E R-2A+M plates, and grown for 3 more days before the start of experiments. If necessary, streptomycin (20 mg ml−1) was added to the medium for selection. Carbon source screen The minimal medium agar plates were pre- pared with a common minimal medium com- position (80). One liter of the final medium contained 2.4 g K2HPO4, 1.08 g NaH2PO4 · 2H2O, 1.62 g NH4Cl and 0.2 g MgSO4 · 7H2O, and 15 g noble agar (Becton, Dickinson and Company). The medium was supplemented with the fol- lowing trace elements: 15 mg Na2EDTA · 2H2O, 3 mg FeSO4 · 7H2O, 4.5 mg ZnSO4 · 7H2O, 3 mg CoCl2 · 6H2O, 0.64 mg MnCl2, 1 mg H3BO3, 0.4 mg Na2MoO4 · 2H2O, 0.3 mg CuSO4 · 5H2O, and 3 mg CaCl2 · 2H2O, and vitamins: 500 mg D-pantothenic acid hemi calcium salt, 100 mg biotin, 400 mg riboflavin, 400 mg thiamine HCl, 200 mg pyridoxal HCl, 150 mg p-amino benzoic acid, 200 mg cobalamin, 50 mg lipoic acid, 150 mg nicotinic acid, and 100 mg folic acid. All media components were prepared with Milli-Q quality water (Millipore). Each carbon source was added as a 10× stock solution to the premixed medium, except for coniferyl alcohol and tyrosine, which were added to the medium directly as powder. The concentration of each carbon source was nor- malized to 30 mM carbon (e.g., 5 mM glucose and 30 mM methanol). A list of all 45 carbon sources and final concentrations can be found in table S1. At-LSPHERE isolates grown on R-2A+M medium were suspended (1 ml loopful) in 1 ml 10 mM magnesium chloride solution, which corresponds to an optical density at 600 nm (OD600) of approximately 0.1 to 0.3. The cell suspensions were transferred to 96-well plates before spotting on agar plates, leaving every other well empty in a checkerboard manner to reduce the risk of contamination and allow more space on the agar plate for each strain. Increasing the distance between spots also strongly reduces the risk of strain interactions based on secondary metabolite production (e.g., antibiotics) that could lead to false-negative results (42). Cell suspensions (2 ml) were spotted on top of each minimal medium plate con- taining individual carbon sources, as well as on a minimal medium plate without carbon source and a R-2A+M plate. Spotting was carried out using a Rainin Liquidator Man- ual Pipetting System (Mettler Toledo). Plates were dried under laminar flow. Bacteria were incubated at 22°C and photographs were taken at 5, 7, and 10 days after spotting. In sum, the screen comprised three total rounds includ- ing 96 strains each, and the last round also comprised 56 randomly selected strains that were screened a second time for validation purposes (table S7). Plate images are availa- ble through Zenodo (81). Analysis of bacterial growth in carbon source screen Colony growth was scored for all strains after 7 days (0 = no growth, 1 = growth) by compar- ing growth on each carbon source with the control plate without a carbon source. We used a dual approach incorporating automated growth scoring based on pixel intensities, using a common threshold for all strains and manual scoring through visual inspection. For the automated analysis, we developed the software tool platescan (github.com/ MicrobiologyETHZ/platescan), which enables unbiased scoring of all strains over different conditions with a firm growth cutoff. Briefly, platescan uses cross-correlation techniques to crop the plate image, locate the colony grid layout, and then determine the best-fitting colony size and location on the basis of the assumption that they are approximately cir- cular. The pixel intensities are rescaled between 1 and 99% of the intensity distribution to en- sure that the images have approximately the same contrast. The program then reports fore- ground and background pixel intensities in red, green, blue, and grayscale, which are then used to threshold growth versus no growth. An example can be found on the program’s GitHub repository. We used platescan to assign growth values to each strain with the following parameters: -r:20, -p: 10,–min_r: 15,–max_r: 40,–edge: 200 40 200 40. In addition, two parameters were set indi- vidually for each screening round because of variation in picture zoom level. Screen1: -x 105 -y 108, screen2: -x 110 -y 109, screen3: -x 114 -y 113. We used the pixel intensities reported in the red channel with a defined minimum thresh- old to consider a strain as having grown. Pixel intensities were obtained by subtracting back- ground values from foreground values for each colony and subsequently subtracting the no-carbon control value. For the few cases in which strains spread into the area used for background intensity calculation, the back- ground value of the neighboring colony was subtracted. For a fraction of strains, we also observed substantial growth without the addition of a carbon source to the medium, suggesting that these strains could either grow on a compo- nent or impurity within the agar or were able to grow on volatile compounds present in the surrounding environment. Therefore, we also manually scored growth for all strains rela- tive to the control plate containing no carbon. Because the magnitude of colony formation differed between carbon sources, assigning a binary growth (1) or no growth (0) value was challenging for strains that exhibited residual growth on the no-carbon control or were gen- erally slow growing. We therefore introduced a third category of nonsignificant (NS) growth for the initial screen analysis but considered them as no growth for subsequent analyses to keep the number of false positives low (fig. S11A). The results of the computational analysis with platescan and the manual analysis showed a very high overlap between the two methods. The highest overlap (with 94.6% of matches com- pared with the manually scored data) was ob- tained with a cutoff of 20 arbitrary units (a.u.). The comparison of the manually and com- putationally scored data showed that the computational method scored growth more often than the manual analysis (n ¼ 485), but most of these were scored as NS initially (n ¼ 332) (fig. S12). To achieve the best accuracy possible, we applied a rigorous curation pro- cess to the screening data. First, we reinspected all cases that did not agree between the manual and computational analysis (20 a.u. cutoff) and all cases that were scored as NS manually. This revealed few false-positive and false-negative scores for both the manual and computational analysis. In the case of the computational analysis, these mistakes were due to either a slight misalignment of the grid or reflections on the plate. To account for slow-growing strains and strains that exhibited high background growth on the no-carbon control condition, we also reinspected all instances that were close to the chosen pixel cutoff (10 to 40 a.u.). For these, we also inspected the growth after 10 days and scored growth on the basis of whether there was an increase in colony density over time. This latter step reduced the number of NS values observed in the initial analysis and mainly improved scoring of the previously mentioned slow-growing Actinobacteria (fig. S11B). The curation procedure reduced the false-positive rate compared with the original analysis from 3.3 to 2.3%, whereas the false-negative rate increased from 5.8 to 6.5%, with most uncer- tainty remaining for Methylobacterium spp. and some members of the Actinobacteria. Plant cultivation Plants were cultivated as described previously (19). In brief, A. thaliana Columbia (Col-0) seeds were surface sterilized (82) and stratified for 4 days at 4°C. Arabidopsis were cultivated in six- well tissue culture plates (TechnoPlasticProducts) filled with 5 ml washed and heat-sterilized calcined clay mixed with 2.5 ml half-strength Murashige & Skoog medium pH 5.8 includ- ing vitamins (½ MS, Duchefa). Seeds were placed in the center of each well. If a seedling did not germinate, a different seedling was transplanted from a separate plate after 7 days. Starting at day 7, each well was supplemented twice per week with 200 ml ½ MS, respectively. Plates were incubated in a growth chamber (Percival, CU41-L4) set to 22°C and 54% relative humidity with an 11-hour photoperiod, fitted with full-spectrum lights (Philips Master TL-D 18 W/950 Graphica) and lights emitting a higher fraction of UVA and UVB (Sylvania Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 9 of 13 RES EARCH | R E S E A R C H A R T I C L E Reptistar F18W/6500 K). The combined light intensity was set to 200 to 210 mmol m−2 s−1 for wavelength 400 to 700 nm and 4 to 5 mmol m−2 s−1 for wavelength 280 to 400 nm. on minimal medium containing isoleucine. Mock treated plants (n ¼ 12) were included in each plant experiment to detect any system- atic contamination with external bacteria. Phyllosphere inoculation Bacteria grown on R-2A+M agar plates were suspended in 10 mM MgCl2 solution, and the OD600 was adjusted to 0.2 for each strain. The final inoculation suspension had an OD600 of 0.02 for all treatments. For single-strain inocu- lations, 150 ml of OD-adjusted strain suspension was added to 1.35 ml 10 mM MgCl2 solution. For two- and three-strain combinations, 75 or 50 ml of each strain were added, respectively. For the seven-strain community, 100 ml of each OD-adjusted strain suspension was mixed, and then a 10-fold dilution was prepared for the final inoculum. The final suspension was mixed well, and then 10-day-old Arabidopsis seedlings were inoculated by slowly pipetting 50 ml over the whole seedling. A 10-fold dilu- tion series was prepared of each inoculum and spotted on R-2A+M agar plates to enumerate total bacterial load. For strain mixes, appro- priate dilutions were plated on R-2A+M agar plates to verify presence of all strains on the basis of colony morphology. Phyllosphere harvest and bacterial colony forming unit (CFU) enumeration Bacterial colonization in the phyllosphere was enumerated when plants were 28 or 29 days old. The whole phyllosphere was harvested with sterile tweezers and a scalpel and was placed in a 2-ml tube containing 200 ml 100 mM phosphate buffer pH 7 and a sterile 5-mm metal ball. The weight of the tube was recorded with and without the plant for plant fresh-weight calculation. The harvested phyllosphere was subsequently crushed for 45 s at 30 Hz with a TissueLyser II (Qiagen). Phosphate buffer (600 ml) was added to the crushed plant material and was mixed thoroughly; 100 ml of this suspension were transferred to a 96-well plate to prepare a 10-fold dilution series in 100 mM phosphate buffer. The dilution series was spotted on R-2A+M agar plates. In addition, 50 ml of each 10−3- and 10−4-fold dilutions were plated on 9-cm round R-2A+M agar plates. If selective plates contain- ing streptomycin were used (for Sphingomonas selection), 50 ml of dilutions 10−1 and 10−2 were plated in addition. Plates were incubated at room temperature and CFUs were counted after 1 to 3 days on dilution series and after 4 to 7 days on round plates. If a strain was not found on the lowest-available dilution, its value was set to 0.9 CFUs for this sample for further anal- ysis. Rhodococcus sp. Leaf233 was selectively grown on minimal medium supplemented with maltose (5 mM) when combined with Pseudomonas sp. Leaf15. Colonies of Leaf233 in mixtures with Microbacterium sp. Leaf179 were differentiated by re-streaking colonies Shake flask cultivations At-LSPHERE isolates grown on R-2A+M me- dium were suspended in 4 ml 10 mM magne- sium chloride solution at an approximate OD600 of 3. Each strain was inoculated at an OD600 of 0.025 into 100-ml baffled Erlenmeyer flasks containing 10 ml of a liquid minimal medium with the same base composition (ions and vitamins) as in the carbon source screen. This medium contained all 44 growth yielding car- bon sources used in the in vitro screen (table S1) at a total concentration of 10 mM C, with each carbon source at a relative concentration corresponding to the medium composition used in the modeling. Strains were inoculated in four biological replicates in monocultures and in the pairwise and community combina- tions used in the plant experiments. Cultures were grown at 22°C with shaking at 200 rpm for 60 hours to allow strains to reach stationary phase. Cell numbers in each culture were enu- merated through a 10-fold dilution series spotted on R-2A+M agar plates. Data analysis Carbon source–utilization data, plant colo- nization experiments, and shake flask experi- ments were analyzed with R 4.04 in RStudio. For the carbon source screen, false-positive and false-negative rates were calculated on the basis of 56 isolates that were screened twice (table S7). The strain phylogeny was based on full-length 16S ribosomal RNA gene sequences extracted from the genome sequence of each strain as described previously (18). Strains that did not grow on any carbon source were ex- cluded from all further analyses. The Manhattan distances of all strains based on carbon source utilization were calculated with the vegdist function of the vegan 2.5 package (83). Hier- archical clustering was conducted with the hclust function with Ward method (ward.D2). Principal coordinate analysis was performed with the cmdscale function in MATLAB R2021a. Bacterial colonization data was log10 trans- formed to calculate the median colonization and statistical significance on the basis of the Wilcoxon rank sum test. P-values were corrected for multiple testing with the Holm method. Effect sizes for the shake flask experiments were calculated with the cohens_d function (var.equal = FALSE, hedges.correction = TRUE) in the rstatix package (84). Log2-fold changes (combination/monoassociation) were calculated on the basis of the median absolute abundance. Data was visualized in R with the package ggplot2 (85) and was further annotated with Adobe Illustrator. For data analysis and visu- alization, the following R packages were used: tidyverse (86), ape 5.4-1 (87), ggridges 0.5.3 (88), shades1.4 (89), phylloR (18), and dendextend 1.14 (90). Generation and curation of genome-scale models The tool CarveMe (58) was first used to generate draft metabolic models based on the assembled genomes of each strain in the At-LSPHERE collection [(53), BioProjects PRJNA297956 and PRJEB47672]. The quality of these genomes was assessed using the tool CheckM (91) (table S8), which reported a mean completeness of 0.993 ± 0.007 and a mean contamination of 0.005 ± 0.007 (mean ± SD, [0,1]). We compared the pre- dicted growth capabilities of each draft model against data from our in vitro carbon source screen and performed the following steps on each draft model: First, we used the tool NICEgame [(62), github.com/EPFL-LCSB/NICEgame] to generate sets of new biosynthetic and transport reactions [drawn from the BiGG database (92)] that would allow the model to produce biomass from each growth-supporting carbon source identified in our in vitro screen. We used this process, which relies on known reaction thermo- dynamic constraints (93, 94), to produce at most 10 alternative sets of reactions for each growth- supporting carbon source. We separately integ- rated each set of reactions into the draft model and tested whether it could produce biomass on a simulated minimal medium (table S6) sup- plemented with each of the 45 carbon sources used in the in vitro screen. We also tested the growth of the model when combinations of re- action sets from up to three carbon sources were integrated, selecting the most accurate model (i.e., the most representative with the fewest false positives) as measured by Matthews cor- ½ Þ= ð relation coefficient TP (cid:2) TN (cid:3) FP (cid:2) FN ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi p (cid:4) Þ ð Þ TN þ FN ð Þ TN þ FP ð Þ TP þ FN ð TP þ FP for further curation. Most models reached this stage having had sets of reactions from only one (111 models) or two (82 models) growth- supporting carbon sources integrated, as add- ing reactions that enable growth on one resource can resolve additional gaps that enable growth on other resources. The carbon sources most often used for gap-filling were maltose, succinate, gluconate, and xylose (being used to gap-fill 40, 26, 23, and 21 models, re- spectively). To correct for remaining false nega- tives (i.e., the model did not grow on a carbon source that supported growth in vitro), we further added reactions from a different model in our collection that did recapitulate growth on the relevant carbon source. A maximum false-positive rate of either 10 carbon sources or half of all true positives for the model, which- ever was smaller, was set to avoid integrating an excessive number of new reactions. False positives for a given carbon source were cor- rected by removing internal reactions rele- vant to its metabolism or by restricting the relevant transport reaction when this was not Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 10 of 13 RES EARCH | R E S E A R C H A R T I C L E successful. Model- and carbon source–specific accuracies are summarized in fig. S5, and re- actions added to the models as part of the curation are enumerated in fig. S4 and table S3. We identified no relationship between model accuracy and either the size (fig. S6B) or the completeness of the underlying genome (fig. S13). A final quality control was performed on each model (59), consisting of testing for mass and charge balance; performing a leak test to ensure no metabolites could be produced from nothing; standardizing the metabolite name- space; adding reaction subsystems; and adding additional gene, metabolite, and reaction iden- tifiers when available. A final report was gen- erated for each model with the validation tool MEMOTE [total score = 0.84 ± 0.02 (mean ± SD, n ¼ 224)] (95), and balanced accuracies for individual models and pairwise interaction pre- Þ=2, dictions were calculated as TPR þ TNR where TPR and TNR represent the true-positive rate and false-positive rates, respectively. All scripts for model generation and simulation, as well as the models and quality reports, are avail- able at github.com/VorholtLab/i-At-LSPHERE. ð Computing mono- and coculture growth All growth simulations were performed using the COBRA Toolbox v2.24.3 (96) with the CPLEX solver v12.10 (IBM) in MATLAB R2021a (MathWorks). Nonlimiting amounts ( vmax ¼ 1000 mmol gDW–1 hour–1) of a minimal me- dium composition containing ions, water, and sources of nitrogen, sulfur, and phosphorus were provided to the models, along with vita- mins at vmax ¼ 0.15 mmol gDW–1 hour–1 (table S6). Limiting amounts of the 45 carbon sources tested in our in vitro screen were provided at abundances intended to broadly estimate the relative availabilities of resource types on leaf surfaces (36, 38, 39, 54) (vmax ¼ 0.15 mmol gDW –1 hour–1 for sugars and organic acids, 0.075 mmol gDW–1 hour–1 for amino acids, and 1.5 mmol gDW–1 hour–1 for methanol). Model growth was simulated with biomass as the objective function and a minimal ATP maintenance flux of 0.5 mmol gDW–1 hour–1. Optimizations were also set to minimize all reaction fluxes to more closely simulate efficient proteome utilization and minimize metabolite cycling (97). For each pair or community, the growth of each strain was first simulated in- dividually, and the resulting biomass flux val- ues and resource uptake fluxes were recorded. Models were then merged by integrating them into a common extracellular compartment (98, 99). Coupling constraints were introduced to avoid infeasible solutions in which one or- ganism produced metabolic flux for the other without producing biomass itself (34, 100). Additionally, metabolite uptake directional- ities were fixed to those observed in monocul- ture to minimize inconsistencies in resource preferences between mono- and coculture. The sum of biomass reactions was set as the objective to be optimized, and exchange reac- tion vmax values were set as equal to those of the models in monoculture, thus simulating an equal abundance of resources between the mono- and coculture conditions. The resulting biomass flux values were recorded and used to calculate interaction scores, which were de- fined as the log2 ratio of biomass flux in co- culture to that in monoculture. Interaction scores were limited to between –5 and 5, with extreme log2-fold changes falling outside this range being set to –5 or 5. RE FERENCES AND NOTES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. A. Apprill, Marine Animal Microbiomes: Toward understanding host–microbiome interactions in a changing ocean. Front. Mar. Sci. 4, 222 (2017). doi: 10.3389/fmars.2017.00222 E. A. Archie, J. Tung, Social behavior and the microbiome. Curr. Opin. Behav. Sci. 6, 28–34 (2015). doi: 10.1016/ j.cobeha.2015.07.008 J. Lloyd-Price et al., Strains, functions and dynamics in the expanded Human Microbiome Project. Nature 550, 61–66 (2017). doi: 10.1038/nature23889; pmid: 28953883 R. Knight et al., The microbiome and human biology. Annu. Rev. Genomics Hum. Genet. 18, 65–86 (2017). doi: 10.1146/ annurev-genom-083115-022438; pmid: 28375652 P. Trivedi, J. E. Leach, S. G. Tringe, T. Sa, B. K. Singh, Plant-microbiome interactions: From community assembly to plant health. Nat. Rev. Microbiol. 18, 607–621 (2020). doi: 10.1038/s41579-020-0412-1; pmid: 32788714 L. S. Bittleston, M. Gralka, G. E. Leventhal, I. Mizrahi, O. X. Cordero, Context-dependent dynamics lead to the assembly of functionally distinct microbial communities. Nat. Commun. 11, 1440 (2020). doi: 10.1038/ s41467-020-15169-0; pmid: 32188849 J. E. Goldford et al., Emergent simplicity in microbial community assembly. Science 361, 469–474 (2018). doi: 10.1126/science.aat1168; pmid: 30072533 S. Estrela et al., Functional attractors in microbial community assembly. Cell Syst. 13, 29–42.e7 (2022). doi: 10.1016/ j.cels.2021.09.011; pmid: 34653368 V. Dubinkina, Y. Fridman, P. P. Pandey, S. Maslov, Multistability and regime shifts in microbial communities explained by competition for essential nutrients. eLife 8, e49720 (2019). doi: 10.7554/eLife.49720; pmid: 31756158 A. R. Pacheco, M. L. Osborne, D. Segrè, Non-additive microbial community responses to environmental complexity. Nat. Commun. 12, 2365 (2021). doi: 10.1038/s41467-021- 22426-3; pmid: 33888697 J. Friedman, L. M. Higgins, J. Gore, Community structure follows simple assembly rules in microbial microcosms. Nat. Ecol. Evol. 1, 109 (2017). doi: 10.1038/s41559-017-0109; pmid: 28812687 12. O. S. Venturelli et al., Deciphering microbial interactions in 13. 14. synthetic human gut microbiome communities. Mol. Syst. Biol. 14, e8157 (2018). doi: 10.15252/msb.20178157; pmid: 29930200 A. L. Gould et al., Microbiome interactions shape host fitness. Proc. Natl. Acad. Sci. U.S.A. 115, E11951–E11960 (2018). doi: 10.1073/pnas.1809349115; pmid: 30510004 T. Camenzind, S. Hättenschwiler, K. K. Treseder, A. Lehmann, M. C. Rillig, Nutrient limitation of soil microbial processes in tropical forests. Ecol. Monogr. 88, 4–21 (2018). doi: 10.1002/ ecm.1279 15. M. E. Hibbing, C. Fuqua, M. R. Parsek, S. B. Peterson, Bacterial competition: Surviving and thriving in the microbial jungle. Nat. Rev. Microbiol. 8, 15–25 (2010). doi: 10.1038/ nrmicro2259; pmid: 19946288 K. Vrede, M. Heldal, S. Norland, G. Bratbak, Elemental composition (C, N, P) and cell volume of exponentially growing and nutrient-limited bacterioplankton. Appl. Environ. Microbiol. 68, 2965–2971 (2002). doi: 10.1128/ AEM.68.6.2965-2971.2002; pmid: 12039756 J. Kehe et al., Positive interactions are common among culturable bacteria. Sci. Adv. 7, eabi7159 (2021). doi: 10.1126/ sciadv.abi7159; pmid: 34739314 16. 17. Nat. Ecol. Evol. 3, 1445–1454 (2019). doi: 10.1038/s41559- 019-0994-z; pmid: 31558832 19. M. Schafer, C. M. Vogel, M. Bortfeld-Miller, M. Mittelviefhaus, J. A. Vorholt, Mapping phyllosphere microbiota interactions in planta to establish genotype-phenotype relationships. Nat. Microbiol. 7, 856–867 (2022). doi: 10.1038/s41564-022- 01132-w; pmid: 35637327 20. K. R. Foster, T. Bell, Competition, not cooperation, dominates interactions among culturable microbial species. Curr. Biol. 22, 1845–1850 (2012). doi: 10.1016/j.cub.2012.08.005; pmid: 22959348 R. Levins, Evolution in Changing Environments: Some Theoretical Explorations (Princeton Univ. Press, 1968). 22. S. H. Hurlbert, The Measurement of Niche Overlap and Some Relatives. Ecology 59, 67–77 (1978). doi: 10.2307/1936632 21. 23. M. Wilson, S. E. Lindow, Coexistence among epiphytic bacterial populations mediated through nutritional resource partitioning. Appl. Environ. Microbiol. 60, 4468–4477 (1994). doi: 10.1128/aem.60.12.4468-4477.1994; pmid: 16349462 24. S. P. Li, J. Tan, X. Yang, C. Ma, L. Jiang, Niche and fitness differences determine invasion success and impact in laboratory bacterial communities. ISME J. 13, 402–412 (2019). doi: 10.1038/s41396-018-0283-x; pmid: 30254322 26. 25. M. Advani, G. Bunin, P. Mehta, Statistical physics of community ecology: A cavity solution to MacArthur’s consumer resource model. J. Stat. Mech. 2018, 033406 (2018). doi: 10.1088/1742-5468/aab04e; pmid: 30636966 E. Nestor, G. Toledano, J. Friedman, Interactions between culturable bacteria are predicted by individual species’ growth. mSystems 8, e0083622 (2023). doi: 10.1128/ msystems.00836-22; pmid: 36815773 L. C. Birch, The meanings of competition. Am. Nat. 91, 5–18 (1957). doi: 10.1086/281957 27. 28. P. Chesson, Mechanisms of maintenance of species diversity. Annu. Rev. Ecol. Syst. 31, 343–366 (2000). doi: 10.1146/ annurev.ecolsys.31.1.343 29. S. Freilich et al., Competitive and cooperative metabolic interactions in bacterial communities. Nat. Commun. 2, 589 (2011). doi: 10.1038/ncomms1597; pmid: 22158444 30. M. Tikhonov, Community-level cohesion without cooperation. 31. eLife 5, e15747 (2016). doi: 10.7554/eLife.15747; pmid: 27310530 J. D. Orth, I. Thiele, B. Ø. O. Palsson, What is flux balance analysis? Nat. Biotechnol. 28, 245–248 (2010). doi: 10.1038/ nbt.1614; pmid: 20212490 32. A. Bordbar, J. M. Monk, Z. A. King, B. O. Palsson, Constraint- based models predict metabolic and associated cellular functions. Nat. Rev. Genet. 15, 107–120 (2014). doi: 10.1038/ nrg3643; pmid: 24430943 33. N. Y. D. Ankrah et al., Enhancing microbiome research through genome-scale metabolic modeling. mSystems 6, e0059921 (2021). doi: 10.1128/mSystems.00599-21; pmid: 34904863 34. S. Magnúsdóttir et al., Generation of genome-scale metabolic reconstructions for 773 members of the human gut microbiota. Nat. Biotechnol. 35, 81–89 (2017). doi: 10.1038/ nbt.3703; pmid: 27893703 35. M. Kumar, B. Ji, K. Zengler, J. Nielsen, Modelling approaches for studying the microbiome. Nat. Microbiol. 4, 1253–1267 (2019). doi: 10.1038/s41564-019-0491-9; pmid: 31337891 37. 38. 36. S. E. Lindow, M. T. Brandl, Microbiology of the phyllosphere. Appl. Environ. Microbiol. 69, 1875–1883 (2003). doi: 10.1128/ AEM.69.4.1875-1883.2003; pmid: 12676659 J. A. Vorholt, Microbial life in the phyllosphere. Nat. Rev. Microbiol. 10, 828–840 (2012). doi: 10.1038/nrmicro2910; pmid: 23154261 J. H. J. Leveau, S. E. Lindow, Appetite of an epiphyte: Quantitative monitoring of bacterial sugar consumption in the phyllosphere. Proc. Natl. Acad. Sci. U.S.A. 98, 3446–3453 (2001). doi: 10.1073/pnas.061629598; pmid: 11248098 F. Ryffel et al., Metabolic footprint of epiphytic bacteria on Arabidopsis thaliana leaves. ISME J. 10, 632–643 (2016). doi: 10.1038/ismej.2015.141; pmid: 26305156 40. G. Innerebner, C. Knief, J. A. Vorholt, Protection of 39. Arabidopsis thaliana against leaf-pathogenic Pseudomonas syringae by Sphingomonas strains in a controlled model system. Appl. Environ. Microbiol. 77, 3202–3210 (2011). doi: 10.1128/AEM.00133-11; pmid: 21421777 41. U. Ritpitakphong et al., The microbiome of the leaf surface of Arabidopsis protects against a fungal pathogen. New Phytol. 210, 1033–1043 (2016). doi: 10.1111/nph.13808; pmid: 26725246 18. C. I. Carlström et al., Synthetic microbiota reveal priority effects and keystone strains in the Arabidopsis phyllosphere. 42. C. M. Vogel, D. B. Potthoff, M. Schäfer, N. Barandun, J. A. Vorholt, Protective role of the Arabidopsis leaf Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 11 of 13 RES EARCH | R E S E A R C H A R T I C L E microbiota against a bacterial pathogen. Nat. Microbiol. 6, 1537–1548 (2021). doi: 10.1038/s41564-021-00997-7; pmid: 34819644 43. C. E. Morris, L. L. Kinkel, “Fifty years of phyllosphere microbiology: significant contributions to research in related fields” in Phyllosphere Microbiology, S. E. Lindow, E. I. Hecht-Poinar, V. J. Elliott, Eds. (APS Press, 2002), pp. 365–375. J. Peñuelas, J. Terradas, The foliar microbiome. Trends Plant Sci. 19, 278–280 (2014). doi: 10.1016/j.tplants.2013.12.007; pmid: 24439491 44. 45. B. K. Singh, P. Trivedi, E. Egidi, C. A. Macdonald, M. Delgado-Baquerizo, Crop microbiome and sustainable agriculture. Nat. Rev. Microbiol. 18, 601–602 (2020). doi: 10.1038/s41579-020-00446-y; pmid: 33037425 F. Martinelli et al., Advanced methods of plant disease detection. A review. Agron. Sustain. Dev. 35, 1–25 (2015). doi: 10.1007/s13593-014-0246-1 46. 47. A. R. Pacheco, J. A. Vorholt, Resolving metabolic interaction mechanisms in plant microbiomes. Curr. Opin. Microbiol. 74, 102317 (2023). doi: 10.1016/j.mib.2023.102317; pmid: 37062173 48. A. M. Geller, A. Levy, “What I cannot create, I do not understand”: Elucidating microbe-microbe interactions to facilitate plant microbiome engineering. Curr. Opin. Microbiol. 72, 102283 (2023). doi: 10.1016/j.mib.2023.102283; pmid: 36868050 J. A. Vorholt, C. Vogel, C. I. Carlström, D. B. Müller, Establishing Causality: Opportunities of synthetic communities for plant microbiome research. Cell Host Microbe 22, 142–155 (2017). doi: 10.1016/j.chom.2017.07.004; pmid: 28799900 49. 51. 50. A. S. Weiss et al., In vitro interaction network of a synthetic gut bacterial community. ISME J. 16, 1095–1109 (2022). doi: 10.1038/s41396-021-01153-z; pmid: 34857933 J. J. Faith, P. P. Ahern, V. K. Ridaura, J. Cheng, J. I. Gordon, Identifying gut microbe-host phenotype relationships using combinatorial communities in gnotobiotic mice. Sci. Transl. Med. 6, 220ra11 (2014). doi: 10.1126/ scitranslmed.3008051; pmid: 24452263 J. Coker et al., A reproducible and tunable synthetic soil microbial community provides new insights into microbial ecology. mSystems 7, e0095122 (2022). doi: 10.1128/ msystems.00951-22; pmid: 36472419 Y. Bai et al., Functional overlap of the Arabidopsis leaf and root microbiota. Nature 528, 364–369 (2015). doi: 10.1038/ nature16192; pmid: 26633631 53. 52. 63. J. A. Valderrama et al., Bacterial degradation of benzoate: Cross-regulation between aerobic and anaerobic pathways. J. Biol. Chem. 287, 10494–10508 (2012). doi: 10.1074/ jbc.M111.309005; pmid: 22303008 64. G. Fuchs, M. Boll, J. Heider, Microbial degradation of aromatic compounds - from one strategy to four. Nat. Rev. Microbiol. 9, 803–816 (2011). doi: 10.1038/nrmicro2652; pmid: 21963803 65. M. Murillo-Roos, H. S. M. Abdullah, M. Debbar, N. Ueberschaar, M. T. Agler, Cross-feeding niches among commensal leaf bacteria are shaped by the interaction of strain-level diversity and resource availability. ISME J. 16, 2280–2289 (2022). doi: 10.1038/s41396-022-01271-2; pmid: 35768644 66. B. Ryback, M. Bortfeld-Miller, J. A. Vorholt, Metabolic 67. adaptation to vitamin auxotrophy by leaf-associated bacteria. ISME J. 16, 2712–2724 (2022). doi: 10.1038/s41396-022- 01303-x; pmid: 35987782 L. Hemmerle et al., Dynamic character displacement among a pair of bacterial phyllosphere commensals in situ. Nat. Commun. 13, 2836 (2022). doi: 10.1038/s41467-022- 30469-3; pmid: 35595740 68. A. P. Goldberg et al., Emerging whole-cell modeling principles and methods. Curr. Opin. Biotechnol. 51, 97–102 (2018). doi: 10.1016/j.copbio.2017.12.013; pmid: 29275251 69. D. B. Bernstein, B. Akkas, M. N. Price, A. P. Arkin, Critical assessment of E. coli genome-scale metabolic model with high-throughput mutant fitness data, bioRxiv 2023.01.05.522875 [Preprint] (2023), https://doi.org/ 10.1101/2023.01.05.522875. 70. C. M. J. Pieterse et al., Induced systemic resistance by 71. 72. beneficial microbes. Annu. Rev. Phytopathol. 52, 347–375 (2014). doi: 10.1146/annurev-phyto-082712-102340; pmid: 24906124 K. Wang, A. Auzane, K. Overmyer, The immunity priming effect of the Arabidopsis phyllosphere resident yeast Protomyces arabidopsidicola strain C29. Front. Microbiol. 13, 956018 (2022). doi: 10.3389/fmicb.2022.956018; pmid: 36118213 E. J. N. Helfrich et al., Bipartite interactions, antibiotic production and biosynthetic potential of the Arabidopsis leaf microbiome. Nat. Microbiol. 3, 909–919 (2018). doi: 10.1038/ s41564-018-0200-0; pmid: 30038309 73. B. A. Maier et al., A general non-self response as part of plant immunity. Nat. Plants 7, 696–705 (2021). doi: 10.1038/ s41477-021-00913-1; pmid: 34007033 74. R. O. Schlechter, M. Miebach, M. N. P. Remus-Emsermann, 54. H. B. Tukey Jr., The leaching of substances from plants. Annu. Rev. Plant Physiol. 21, 305–324 (1970). doi: 10.1146/ annurev.pp.21.060170.001513 55. M. J. Poupin, T. Ledger, R. Roselló-Móra, B. González, The 75. Arabidopsis holobiont: A (re)source of insights to understand the amazing world of plant-microbe interactions. Environ. Microbiome 18, 9 (2023). doi: 10.1186/s40793-023-00466-0; pmid: 36803555 56. P. Vaz Jauri, L. L. Kinkel, Nutrient overlap, genetic relatedness and spatial origin influence interaction-mediated shifts in inhibitory phenotype among Streptomyces spp. FEMS Microbiol. Ecol. 90, 264–275 (2014). doi: 10.1111/ 1574-6941.12389; pmid: 25098381 57. M. Michalska-Smith et al., Network structure of resource use and niche overlap within the endophytic microbiome. ISME J. 16, 435–446 (2022). doi: 10.1038/s41396-021-01080-z; pmid: 34413476 58. D. Machado, S. Andrejev, M. Tramontano, K. R. Patil, Fast automated reconstruction of genome-scale metabolic models for microbial species and communities. Nucleic Acids Res. 46, 7542–7553 (2018). doi: 10.1093/nar/gky537; pmid: 30192979 I. Thiele, B. O. Palsson, A protocol for generating a high-quality genome-scale metabolic reconstruction. Nat. Protoc. 5, 93–121 (2010). doi: 10.1038/nprot.2009.203; pmid: 20057383 60. S. N. Mendoza, B. G. Olivier, D. Molenaar, B. Teusink, 59. A systematic assessment of current genome-scale metabolic reconstruction tools. Genome Biol. 20, 158 (2019). doi: 10.1186/s13059-019-1769-1; pmid: 31391098 62. 61. D. B. Bernstein, S. Sulheim, E. Almaas, D. Segrè, Addressing uncertainty in genome-scale metabolic model reconstruction and analysis. Genome Biol. 22, 64 (2021). doi: 10.1186/ s13059-021-02289-z; pmid: 33602294 E. Vayena et al., A workflow for annotating the knowledge gaps in metabolic reconstructions using known and hypothetical reactions. Proc. Natl. Acad. Sci. U.S.A. 119, e2211197119 (2022). doi: 10.1073/pnas.2211197119; pmid: 36343249 Driving factors of epiphytic bacterial communities: A review. J. Adv. Res. 19, 57–65 (2019). doi: 10.1016/ j.jare.2019.03.003; pmid: 31341670 T. Orevi, N. Kashtan, Life in a droplet: Microbial ecology in microscopic surface wetness. Front. Microbiol. 12, 655459 (2021). doi: 10.3389/fmicb.2021.655459; pmid: 33927707 J. A. Cole, L. Kohler, J. Hedhli, Z. Luthey-Schulten, Spatially- resolved metabolic cooperativity within dense bacterial colonies. BMC Syst. Biol. 9, 15 (2015). doi: 10.1186/s12918- 015-0155-1; pmid: 25890263 E. Bauer, J. Zimmermann, F. Baldini, I. Thiele, C. Kaleta, BacArena: Individual-based metabolic modeling of heterogeneous microbes in complex communities. PLOS Comput. Biol. 13, e1005544 (2017). doi: 10.1371/ journal.pcbi.1005544; pmid: 28531184 I. Dukovski et al., A metabolic modeling platform for the computation of microbial ecosystems in time and space (COMETS). Nat. Protoc. 16, 5030–5082 (2021). doi: 10.1038/ s41596-021-00593-3; pmid: 34635859 L. Angeles-Martinez, V. Hatzimanikatis, Spatio-temporal modeling of the crowding conditions and metabolic variability in microbial communities. PLOS Comput. Biol. 17, e1009140 (2021). doi: 10.1371/journal.pcbi.1009140; pmid: 34292935 76. 77. 78. 79. 80. R. Peyraud et al., Demonstration of the ethylmalonyl-CoA pathway by using 13C metabolomics. Proc. Natl. Acad. Sci. U.S.A. 106, 4846–4851 (2009). doi: 10.1073/pnas.0810932106; pmid: 19261854 81. M. Schäfer, Images of colony growth for 224 Arabidopsis thaliana bacterial strains on 46 carbon sources, Zenodo (2023); https://doi.org/10.5281/zenodo.7915606. 82. B. Schlesier, F. Bréton, H.-P. Mock, A hydroponic culture system for growing Arabidopsis thaliana plantlets under sterile conditions. Plant Mol. Biol. Report. 21, 449–456 (2003). doi: 10.1007/BF02772594 J. Oksanen et al., vegan: Community Ecology Package, Version 2.6-4, 2022; https://cran.r-project.org/package=vegan. 83. 84. A. Kassambara, rstatix: Pipe-Friendly Framework for Basic Statistical Tests, Version 0.7.2, 2023; cran.r-project.org/ web/packages/rstatix. 85. H. Wickham, ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag, 2016); https://ggplot2.tidyverse.org. 86. H. Wickham et al., Welcome to the Tidyverse. J. Open Source 87. Softw. 4, 1686 (2019). doi: 10.21105/joss.01686 E. Paradis, K. Schliep, ape 5.0: An environment for modern phylogenetics and evolutionary analyses in R. Bioinformatics 35, 526–528 (2019). doi: 10.1093/bioinformatics/bty633; pmid: 30016406 88. C. O. Wilke, ggridges: Ridgeline Plots in “ggplot2,” R package Version 0.5.4, 2022; https://wilkelab.org/ggridges/. J. Clayden, shades: Simple colour manipulation, Version 1.4.0, 2019; https://CRAN.R-project.org/package=shades. 89. 90. T. Galili, dendextend: An R package for visualizing, adjusting and comparing trees of hierarchical clustering. Bioinformatics 31, 3718–3720 (2015). doi: 10.1093/bioinformatics/btv428; pmid: 26209431 91. D. H. Parks, M. Imelfort, C. T. Skennerton, P. Hugenholtz, G. W. Tyson, CheckM: Assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 25, 1043–1055 (2015). doi: 10.1101/gr.186072.114; pmid: 25977477 92. Z. A. King et al., BiGG Models: A platform for integrating, standardizing and sharing genome-scale models. Nucleic Acids Res. 44 (D1), D515–D522 (2016). doi: 10.1093/ nar/gkv1049; pmid: 26476456 93. C. S. Henry, L. J. Broadbelt, V. Hatzimanikatis, Thermodynamics-based metabolic flux analysis. Biophys. J. 92, 1792–1805 (2007). doi: 10.1529/biophysj.106.093138; pmid: 17172310 94. P. Salvy et al., pyTFA and matTFA: A Python package and a Matlab toolbox for Thermodynamics-based Flux Analysis. Bioinformatics 35, 167–169 (2019). doi: 10.1093/ bioinformatics/bty499; pmid: 30561545 95. C. Lieven et al., MEMOTE for standardized genome-scale 96. metabolic model testing. Nat. Biotechnol. 38, 272–276 (2020). doi: 10.1038/s41587-020-0446-y; pmid: 32123384 L. Heirendt et al., Creation and analysis of biochemical constraint-based models using the COBRA Toolbox v.3.0. Nat. Protoc. 14, 639–702 (2019). doi: 10.1038/ s41596-018-0098-2; pmid: 30787451 97. N. E. Lewis et al., Omic data from evolved E. coli are consistent with computed optimal growth from genome-scale models. Mol. Syst. Biol. 6, 390 (2010). doi: 10.1038/ msb.2010.47; pmid: 20664636 98. N. Klitgord, D. Segrè, Environments that induce synthetic microbial ecosystems. PLOS Comput. Biol. 6, e1001002 (2010). doi: 10.1371/journal.pcbi.1001002; pmid: 21124952 99. A. Heinken, I. Thiele, Anoxic conditions promote species- specific mutualism between gut microbes in silico. Appl. Environ. Microbiol. 81, 4049–4061 (2015). doi: 10.1128/ AEM.00101-15; pmid: 25841013 100. A. Heinken, S. Sahoo, R. M. T. Fleming, I. Thiele, Systems- level characterization of a host-microbe metabolic symbiosis in the mammalian gut. Gut Microbes 4, 28–40 (2013). doi: 10.4161/gmic.22370; pmid: 23022739 101. A. R. Pacheco, VorholtLab/i-At-LSPHERE: v1.0, Zenodo (2023); https://doi.org/10.5281/zenodo.7915630. 102. C. M. Field, MicrobiologyETHZ/platescan: v1.0, Zenodo (2023); https://doi.org/10.5281/zenodo.7898736. AC KNOWLED GME NTS We thank S. Hegi and M. Berger for experimental support; A. Şahin, O. Oftadeh, and D. Machado for helpful discussions on the metabolic modeling; and C. Pestalozzi for contributing to the visualization shown in Fig. 3B. Funding: This study was supported by Swiss National Science Foundation grant NRP72 (J.A.V.), a European Research Council Advanced Grant (PhyMo, no. 668991) (J.A.V.), Swiss National Science Foundation grant 200021_188623 (V.H.), NCCR Microbiomes, Swiss National Science Foundation (51NF40_180575) (V.H. and J.A.V.), and a James S. McDonnell Postdoctoral Fellowship (2020-1332) (A.R.P.). Author contributions: Conceptualization: M.S., A.R.P., and J.A.V. Methodology: M.S., A.R.P., C.M.F., E.V., and V.H. Investigation: M.S., A.R.P., R.K., and M.B.-M. Visualization: M.S. and A.R.P. Funding acquisition: A.R.P., V.H., and J.A.V. Project administration: J.A.V. Supervision: V.H. and J.A.V. Writing – original draft: M.S., A.R.P., and J.A.V. Writing – review and editing: M.S., A.R.P., R.K., M.B.-M., E.V., C.M.F., V.H., and J.A.V. Competing interests: The authors declare that they have no competing Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 12 of 13 RES EARCH | R E S E A R C H A R T I C L E interests. Data and materials availability: Materials and methods are available as supplementary materials. The collection of genome-scale models, as well as scripts for generating the models and conducting simulations are available through Zenodo (101). The platescan software (102) and plate images used for the carbon source screen (81) are available through Zenodo. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse Tables S1 to S8 MDAR Reproducibility Checklist SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf5121 Figs. S1 to S13 View/request a protocol for this paper from Bio-protocol. Submitted 27 October 2022; resubmitted 23 March 2023 Accepted 18 May 2023 10.1126/science.adf5121 Schäfer et al., Science 381, eadf5121 (2023) 7 July 2023 13 of 13
10.1126_science.adg0727
RES EARCH MICROBIOTA Emergent coexistence in multispecies microbial communities Chang-Yu Chang1,2,3*, Djordje Bajić1,2,4*, Jean C. C. Vila1,2, Sylvie Estrela1,2†, Alvaro Sanchez1,2,5* Understanding the mechanisms that maintain microbial biodiversity is a critical aspiration in ecology. Past work on microbial coexistence has largely focused on species pairs, but it is unclear whether pairwise coexistence in isolation is required for coexistence in a multispecies community. To address this question, we conducted hundreds of pairwise competition experiments among the stably coexisting members of 12 different enrichment communities in vitro. To determine the outcomes of these experiments, we developed an automated image analysis pipeline to quantify species abundances. We found that competitive exclusion was the most common outcome, and it was strongly hierarchical and transitive. Because many species that coexist within a stable multispecies community fail to coexist in pairwise co-culture under identical conditions, we concluded that multispecies coexistence is an emergent phenomenon. This work highlights the importance of community context for understanding the origins of coexistence in complex ecosystems. E xplaining species coexistence and the be- wildering diversity of ecological com- munities is a major goal of ecology (1). Historically, this problem has been in- vestigated through the lens of species interactions and population dynamics. This work has played a central role in theoretical ecology (2, 3), establishing, for example, the importance of competitive interactions for community stability (4, 5) and the criteria re- quired for stable coexistence in species pairs and pairwise networks (6, 7). An important caveat is that the ability of any model to fully capture the population dynamics of empirical populations is limited, and interactions be- tween species are often modulated by envi- ronmental context (8, 9) and by the presence of additional species (10, 11). As a result, in recent years, research has started to shift to directly study coexistence networks (12–14). A central question is whether the known out- come of competition between pairs of species, i.e., their coexistence or competitive exclusion, can be leveraged to predict the composition of complex communities and the paths leading to their assembly (12, 13, 15). If this approach were fruitful, then it would circumvent the need to know the full mathematical structure of population dynamics models to predict com- munity assembly (14). There are two opposing views on species coexistence (Fig. 1A). A reductionist perspec- tive is that multispecies coexistence is an ad- 1Department of Ecology & Evolutionary Biology, Yale University, New Haven, CT, USA. 2Microbial Sciences Institute, Yale University, New Haven, CT, USA. 3Department of Biology, University of Pennsylvania, Philadelphia, PA, USA. 4Department of Biotechnology, Delft University of Technology, Delft, Netherlands. 5Department of Microbial Biotechnology. Centro Nacional de Biotecnología - CSIC, Campus de Cantoblanco, Madrid, Spain. *Corresponding author. Email: alvaro.sanchez@cnb.csic.es (A.S.); d.bajic@tudelft.nl (D.B.); cychang2@sas.upenn.edu (C.-Y.C.) †Present address: Department of Bioengineering, Stanford University, Stanford, CA, USA. ditive affair, and all of the coexisting members of a community must also coexist as pairs when isolated from the community context (14). An alternative view is that coexistence in a multispecies community is a more com- plex, or emergent, property of the community, which is not exhibited by its most elementary units of coexistence, pairs of species in isola- tion (16). Which of these two views best reflects the reality of empirical communities (Fig. 1A)? Determining which view is more accurate re- quires deconstructing a community into spe- cies pairs to determine whether all possible combinations can coexist. If most can, then the reductionist view is supported. If few can, then coexistence is an emergent property of the community, as is seen in nontransitive competition [i.e., as in the rock, paper, scissors game (17–24)], which may allow multiple spe- cies to coexist even when none of them do as a pair in isolation (17–24). Resolving this question is essential in mi- crobial ecology given the enormous and still largely unexplained diversity of microbial eco- systems (13, 25). Directly testing the two hypothesized scenarios described above is generally not feasible in natural microbial communities because of their diversity. Even if we managed to isolate most community mem- bers from a given habitat, the number of rep- licate environments that we would need to recreate to culture every single pair would scale quadratically with the community rich- ness. Recent studies have taken a synthetic approach by reconstituting species pairs from natural communities in well-controlled lab- oratory environments (14, 26, 27). Although these studies found support for the reduc- tionist hypothesis, their limitation lies in the small fraction of coexisting species that could be isolated and the differences between the laboratory environment and the original com- munity habitat. Here, we sought to directly test the two hypotheses in an empirical system that is well suited for this purpose. Our starting point was a collection of bacterial enrichment commu- nities that we have recently assembled in well- controlled synthetic environments containing glucose as the single externally supplied limit- ing nutrient (Fig. 1B) (9, 28–31). These com- munities formed in a manner that is similar to the “random zoo” model in theoretical ecology (32). In brief, 12 soil and plant microbiomes were resuspended in separate test tubes con- taining M9 minimal medium (9) (Fig. 1B). This provided us with a diverse pool of bacterial species containing between 110 and 1290 exact sequence variants (ESVs) (fig. S1) (9). These 12 initial microbiota solutions were then in- oculated by a 125-fold dilution into separate bioreactors containing M9-glucose growth medium (see the materials and methods), in- cubated for 48 hours under static conditions at 30°C, and then serially passaged 12 times each (~84 bacterial generations under our conditions) (Fig. 1B) (9). Community composi- tion at various time points was determined by 16S ribosomal RNA (rRNA) amplicon sequenc- ing. All communities contained multiple (N < 25) coexisting ESVs belonging primarily to the families Enterobacteriaceae and Pseudomona- daceae (Fig. 1B) (9, 28, 30, 33). It was thus pos- sible in this system to deconstruct multiple stable communities and reconstitute and com- pete most pairs of species under the same starting conditions. This experimental system allowed us to evaluate whether all pairs of organisms that coexist as a part of a multispe- cies community also coexist in isolation (29), thus directly testing whether coexistence is a pairwise or an emergent phenomenon. RESULTS Coexistence is stable in our enrichment communities To establish whether coexistence in these mul- tispecies enrichment communities is stable, we set out to analyze the published commu- nity assembly dynamics data from previous studies (9, 34, 35) in which the frequencies (xi) of all ESVs were quantified at the end of each transfer (i) for 26 representative com- munities (fig. S2). We determined the inva- sion fitness [F = log(xi/xi–1)] of every ESV in these communities over their full assembly dynamics (i = 2,3, …, 12; see the materials and methods) and found that a large majority of the ESVs found at the end of the experiment exhibited hallmarks of negative frequency– dependent selection. For 95/99 of these ESVs, the dependence between fitness and frequen- cy was best fit by a negative regression slope (fig. S3), and the equilibrium frequency (x*) predicted from this linear regression model [the frequency for which F(x*) = 0] agreed very well with the empirically observed equilibrium Chang et al., Science 381, 343–348 (2023) 21 July 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E A Is species coexistence an emergent property of communities? Community (system) Species pairs (elementary units) No, species coexistence is predictable from pairs Yes, species coexistence is an emergent property Coexistence Exclusion ? Both the community and all its pairs coexist The community coexists but not all the pairs do ESV equilibrium frequency four representative communities in current study other 22 communities C y c n e u q e r f m o r f d e t c d e r i P ) * x ( n o i t c e e s l t n e d n e p e d 1.00 0.75 0.50 0.25 0.00 B (I) Community assembly dynamics 1.00 Colonization Isolation Dilution x12 Growth Stable community Replicate identical abiotic environment Pairs reconstitution Dilution x8 Growth 0.00 0.25 0.50 0.75 1.00 Average frequency over the last four transfers D Pseudomonas.10 x*=0.173 Klebsiella x*=0.762 y c n e u q e r F 0.75 0.50 0.25 2 0 2 0 s s e n t i f i n o s a v n I 0.00 inoculum 1 2 3 4 5 6 7 8 9 10 11 12 Transfer 0.00 0.25 0.50 0.75 1.00 Frequency (II) Isolated abundant species y c n e u q e r F 1.00 0.75 0.50 0.25 0.00 1 2 3 4 5 6 7 8 9 10 11 12 Community Enterobacteriaceae Citrobacter Enterobacteriaceae Klebsiella Raoultella Raoultella.1 Salmonella Salmonella.1 Yersinia Pseudomonadaceae Pseudomonas Pseudomonas.1 Pseudomonas.10 Pseudomonas.13 Pseudomonas.2 Pseudomonas.3 Pseudomonas.4 Pseudomonas.9 Comamonadaceae Comamonadaceae Comamonadaceae.1 Comamonas.1 Delftia Delftia.1 Aeromonadaceae Aeromonas Moraxellaceae Acinetobacter.2 Acinetobacter.3 Alcaligenaceae Bordetella.1 Other Not isolated Fig. 1. Enrichment microbial communities allowed us to test the complexity of species coexistence. (A) The two hypotheses about species coexistence tested in our study. (B) To discriminate between the two hypotheses, we used an empirical system constructed from previously assembled enrichment in vitro bacterial communities under serial growth and dilution cycles (9). In inset I, we present the full assembly dynamics for a representative community, showing the frequency of each ESV at the end of every growth period (transfers). We only show ESVs >2% in frequency, each in a different color. We chose 12 representative communities with richness ranging between N = 5 and N = 13 ESVs at transfer 12 (inset II) and isolated most community members (colored bars) covering an average of 89.4% of the abundance. Gray bars represent ESVs that we were not able to isolate (see the materials and methods). Raw data were obtained from previous studies (9, 34, 35). (C) Frequency-dependent dynamics predicted the empirically observed equilibrium frequencies. Empirical equilibrium frequencies (horizontal axis) were quantified as the average frequency of an ESV in the last four transfers of the community assembly process (transfers nine to 12). To determine the predicted equilibrium frequency x* (x axis), we first quantified the invasion fitness Fi = log (xi/xi–1) for each ESV at each transfer and then regressed this Fi against ESV frequency. This regression yielded a negative slope for 95/99 ESVs found near the equilibrium in their respective community (fig. S3), indicating that these ESVs are subject to negative frequency–dependent selection. In these cases, we estimated the equilibrium frequency x* as the x-intercept of the regression line (figs. S3 and S4). (D) Two examples of invasion fitness analysis from the community in inset I showing negative frequency–dependent selection. The yellow line represents the linear fit as determined by least-squares regression (N = 11, R2 = 0.92 and N = 11, R2 = 0.70 for the top and bottom panels, respectively). The x-intercept was used to estimate the equilibrium frequency x*, which is shown as a vertical dashed line. frequencies, which we determined as the av- erage frequency of the ESV over the last four transfers (Fig. 1C, fig. S3, and materials and methods). By contrast, ESVs that were only tran- siently present during community assembly but were not part of the final stable community generally exhibited either negative average fitness values or equilibrium frequencies close to 0 (figs. S4 and S5). Overall, our quantitative analyses indicated that the ESVs that were pres- ent in the final transfer of our multispecies enrichment communities could invade from low frequency, fulfilling the mutual invisibility criterion of stable coexistence (36). Quantification of pairwise competition assays To empirically test whether stable multispe- cies coexistence was a pairwise phenomenon in our enrichment communities, we chose 12 representative communities containing between five and 13 ESVs in stable equilibrium, plated them on their final transfer, and then selected at least three morphologically distinct isolates from each community (fig. S6 and materials and methods). Using Sanger sequencing, we obtained the full-length sequence of the 16S rRNA gene of these isolates, aligned it with the ESVs that were found in their communities of origin, and retained all isolates with at least 200–base pair consensus sequence and four or fewer mismatches. This resulted in a total of 62 isolates, 40 with fully matching align- ments and 22 with one to four mismatches Chang et al., Science 381, 343–348 (2023) 21 July 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E A Pairwise competition experiment Initial frequencies One stable community Isolation 95% 5% 50% 50% 5% 95% Dilution x8 Growth Start End B n o i t c a r F 1.0 0.8 0.6 0.4 0.2 0.0 6 3 2 9 3 333 7 3 33 6 12 5 4 4 4 6666 666 666 9 5 888 9 13 11 7 8 5 9 10 7 5 9999 101010 2121212121 3030303030 4040404040 n. of ESVs n. of isolates n. of tested pairs competitive exclusion on the path to competitive exclusion stable coexistence (mutual invasibility) coexistence without evidence of mutual invasibility inconclusive 9 10 11 12 1 2 3 4 6 7 5 8 Community C y c n e u q e r F 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 Competitive exclusion On the path to competitive exclusion Stable coexistence (mutual invasibility) Coexistence without evidence of mutual invasibility Inconclusive 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 y c n e u q e r F 0 8 0 8 y c n e u q e r F 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 y c n e u q e r F 0 8 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 y c n e u q e r F 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 0 8 0 8 0 8 8 0 Transfer 0 8 0 8 0 8 0 8 0 8 8 0 Transfer 0 8 0 8 0 8 0 8 8 0 Transfer 8 0 Transfer 0 8 Transfer Initial frequencies 95% 50% 5% Fig. 2. Multispecies coexistence is an emergent property of the community. (A) To determine whether isolated species pairs coexist or outcompete one another, we cultured each pair at three different initial frequencies. Pairs were propagated in the same culture conditions as their community of origin for eight consecutive passages. The pairwise competition outcomes of all 12 enrichment communities are shown in (B), and communities are ordered by the number of strains in each community from the smallest (three taxa) to the largest (10 taxa). The numbers above in each bar show the number of ESVs, the number of isolated strains, and the number of tested pairs, respectively. Note that some communities have missing pairs because these pairs either did not have any colonies in co-culture or had low classification model accuracy. (C) Competition outcomes of 144 pairwise co-cultures. Mean frequencies and 95% confidence intervals were determined by Poisson sampling (N = 1000; see the materials and methods). For clarity, we plotted in all cases the frequency of the isolate ending with a lower average frequency in time point 8 (T8). In coexisting pairs, the mean equilibrium frequency on the final transfer is represented by a horizontal dashed line, and the 95% confidence interval (computed from Poisson sampling, N = 1000) as a shaded area around it. Each of the inset grids indicates the change in frequency from the initial time point (T0) to the final one (T8). The background color represents the competition outcomes, and the line color indicates the three initial frequencies. To establish significant changes in frequency between T0 and T8 in each experiment, we used Wilcoxon–Mann-Whitney tests with N = 2000 and a significance threshold of P < 0.05 (see the materials and methods). (fig. S7 and materials and methods), covering on average 89.4% of the ESV composition of the original communities (Fig. 1B). We then performed every possible pairwise competition experiment among the isolates of each community by mixing inocula of pairs of isolates and passaging each mixture for eight growth-dilution cycles in the same glucose minimal medium at the same temperature (30°C) used in the original community enrich- ment experiments (Fig. 2A). All pairwise com- petition experiments were performed three times, each at a different starting count pro- portion of ~5:95, ~50:50, and ~95:5 (Fig. 2A and materials and methods). During each growth cycle, the cells were incubated for 48 hours, after which the resulting culture was diluted 125-fold into fresh medium, as was done in the original community assembly experiment (9). At the end of the last dilution cycle, we measured the composition of our pairwise co-cultures by plating them on Petri dishes and counting the colonies belonging to each isolate. To avoid human bias in colony morphology identification, we adopted an automated image- processing pipeline (fig. S8) combined with a machine-learning approach for classification using 159 × 3 = 477 co-culture images on the basis of 40 colony morphology features (figs. S9 and S10, table S1, and supplementary mate- rials). The pipeline started by extracting color channels and correcting for uneven backgrounds, followed by segmenting colony objects and ex- tracting the morphological features from these. These colony features were analyzed using ran- dom forest classification to determine whether each colony present in the co-culture image belonged to one morphotype or another (fig. S10). This approach allowed us to quantify the number of colony-forming units of each of the two competitors in pairwise co-culture. Of Chang et al., Science 381, 343–348 (2023) 21 July 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E A Community 1 2 3 4 5 6 7 8 9 10 11 12 1 2 3 4 5 6 7 8 9 R a n k competitive exclusion on the path to competitive exclusion stable coexistence (mutual invasibility) coexistence without evidence of mutual invasibility B t n u o C 900 600 300 0 0 Number of expected nontransitivity 10 20 30 Fig. 3. Competitive hierarchy prevails among species pairs in stably coexist- ing communities. (A) All isolates in our 12 communities were rank-ordered from top to bottom on the basis of the number of other isolates that they excluded in pairwise competition, using data from the experiments shown in Fig. 2. The gray nodes in the network represent each individual isolate. Red arrows point from the winning isolate of a pairwise competition to the losing one. Blue lines connect isolates that coexist. (B) Binomial distribution with N = 77, P = 0.25 showing the expected number of pairs that exhibit transitivity if we randomly swapped the coexistence and exclusion links in (A). The red open circle marks the experimentally determined number of trios that broke transitivity (in our case, zero). the 159 competing pairs, six did not yield a measurable optical density in either of the three competition assays regardless of their inoculation frequencies, and no colonies were detected. Because we could assign neither co- existence nor competitive exclusion to these pairs, which were also formed by pairs of isolates that were not present at the final transfer in monoculture, we excluded them from further analysis. We removed nine ad- ditional pairs for which the trained model performed poorly on the validation datasets (accuracy score <0.9; fig. S11 and materials and methods). We therefore used N = 144 pairs in our analysis. The automated pipe- line approach agreed well with visual colony identification, yielding comparable results for both the total colony count on a plate [R2 = 0.85; root-mean-square deviation (RMSD) = 17.67; N = 381] and the relative frequency of different colony morphotypes (R2 = 0.87; RMSD = 0.17; N = 381) (fig. S12) for the 127 pairs with an accuracy score > 0.9 that could be discriminated by eye. Multispecies coexistence is an emergent property In 26.4% of the pairs (38/144), one of the two competitors had become competitively excluded by the end of the last dilution cycle in all three competition experiments (i.e., no colonies were detected on the plates) regardless of its starting inoculation proportion (Fig. 2, B and C, dark red). We marked these outcomes as competitive exclusion. For 45.1% of the pairs (65/144), the frequency of the losing species declined (DF < 0) in all three competition experiments regardless of its initial propor- tion (Fig. 2, B and C, light red box, and materials and methods). This indicates that its trajectory was on the path to competitive exclusion. Adding these outcomes to the com- petitive exclusion category, we found that 71.6% of the pairs (103/144) failed to coexist in the absence of the other community mem- bers. These results were not driven by the poor competitive ability of the least-abundant ESVs, because eliminating from the analysis those isolates with ESVs with <0.05 frequency in the stable multispecies communities still produced a majority of competitive exclusion outcomes (61/84 = 72.6%) (fig. S13). All 12 communities contained at least one pair, but generally more, that could not coexist in isolation (Fig. 2B). The fraction of pairs ex- hibiting competitive exclusion was similar across communities regardless of their rich- ness (Fig. 2B). These results are not consistent with the additive assembly rule proposed pre- viously (14), which would have predicted less- diverse communities composed only of those taxa that can coexist in isolated pairs. There- fore, complex multispecies coexistence could not be reduced to pairwise relationships in our communities, and it is thus likely an emergent property of the whole community. A substantial fraction of pairwise competi- tions (28.5% of the pairs, 41/144) did not result in competitive exclusion, indicating that pair- wise coexistence may still be common among members of a stable multispecies community. Among these 41 pairs, 29 were still coexisting in all three competition experiments after eight transfers (Fig. 2, B and C, blue). To identify those pairs that coexist stably, we apply the mutual invasibility criterion, which requires that both species must be able to invade each other from low frequency (36) (Fig. 2, B and C, dark blue). Methodologically, this requires that sign(Dx) = sign[x* – x(T0)] for both species in all three pairwise competition experiments (see the materials and methods). Here, Dx de- notes the change in a species frequency be- tween the final and initial transfers, x* is the equilibrium frequency for that species (which we determined by averaging the final transfer frequencies of the three experiments; see the materials and methods), and x(T0) is the spe- cies’ inoculation frequency on the first day of the experiment. This condition was met in 21 of the 29 coexisting pairs. The criteria for mu- tual invasibility were not met in the remaining eight coexisting pairs, so we classified these as Chang et al., Science 381, 343–348 (2023) 21 July 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E coexisting without evidence of mutual invasi- bility (Fig. 2, B and C, light blue). The remain- ing fraction of our pairs (12/144, 8.3%) did not offer conclusive results because the outcome of the competition was not consistent in the three experiments. We left these as inconclu- sive (Fig. 2, B and C, gray). Competitive exclusion is hierarchical and transitive In an effort to better explore the structure of our pairwise competition network, we used the competition outcomes shown in Fig. 2 to rank all isolates in each community by the number of competitors that each of them ex- cluded (see the materials and methods). We found that competitive exclusion was almost fully hierarchical: In all but one of the 103 pairs in which one of the two isolates was excluded, the lower-rank species was the one that was excluded (Fig. 3A). The ranks in the competi- tive hierarchy were positively correlated with the frequency rank of the corresponding ESV in the parent community (Spearman’s r = 0.42, P < 0.001, N = 62; fig. S14), but this pattern was mostly driven by less-diverse communities (fig. S14), which recapitulates previous findings from plant communities (37). We also found that competitive hierarchy was positively correlated with the strain’s growth rate in glucose medium (Pearson’s r = –0.314, P = 0.0129, N = 62; fig. S15). Regarding the type of metabolism, respiro- fermenters had a higher average competitive rank (mean = 2.54) than obligate respirers (mean = 4.72) (Wilcoxon–Mann-Whitney test P < 0.001, N = 62; fig. S16), a pattern consistent with our previous work (9, 28). An extreme case of emergent coexistence may occur when coexistence networks are non- transitive (17, 24). However, we found that nontransitive cycles were unlikely to stabilize coexistence in our communities. Of 77 triplets of species that could be connected by compet- itive exclusion links in our 12 communities, we did not find a single violation of transitivity (Fig. 3B). Because the expected fraction of nontransitive triplets in a random network is P = 1/4, the probability of observing this out- come by chance is given by P(0) = (1/4)77 = 4.4 × 10−47 (Fig. 3B). Discussion The aim of this study was to empirically test whether coexistence in microbial communi- ties is a pairwise phenomenon or if it is an emergent property of the community. To ad- dress this question, we isolated most mem- bers of 12 stable enrichment communities and determined whether each possible pair could coexist in the absence of the other mem- bers of their communities under the same cul- ture conditions as in the enrichment. Although a substantial fraction of pairs did coexist (29/144, 20.1%), a majority (103/144, 71.5%) of them ended up in competitive exclusion, with one of the two members becoming ex- cluded or on the path to it. This indicates that coexistence could not be reduced to a pairwise phenomenon in our enrichment communities and that the community context is generally required for species pairs to coexist. Our find- ing contrasts with the outcome of a recent empirical study supporting the reductionist hypothesis, which concluded that the coex- istence of multiple species in bottom-up as- sembled communities requires every pair to coexist in isolation (14). Given that both hypotheses can be correct in different communities (14, 16), our results prompt the question of under which condi- tions each is most likely to occur. We have not yet determined whether the complex nature of multispecies coexistence in our enrichment communities derives from higher-order inter- actions, or if it can be explained by a complex network of pairwise interactions. Another pos- sible factor that may stabilize coexistence, but which our study has not addressed, is the rapid emergence of intra-strain diversity through evo- lutionary processes. Evolution of new species interactions, such as the appearance of a new mutualism, may mediate the emergent coex- istence of pairs of strains that would other- wise end in competitive exclusion (16). As for broader evolutionary patterns, we did not find a correlation between pairwise coexistence and sequence similarity (fig. S17), although our analysis was limited to the 16S marker gene. Finally, spatial structure is also known to af- fect microbial coexistence [e.g., (38)], but the number and nature of spatial niches could not be identified in a straightforward manner in our experiments. Theoretical studies have suggested that nontransitivity can stabilize the coexistence of multiple competing species in the presence of spatial heterogeneity (18) or when com- petitors have differential competitive abil- ities on multiple limiting resources (17, 21). Although the idea of nontransitivity is well established in theory, empirical studies on its prevalence are sparse. Our mutual inva- sion experiments with 144 species pairs from each of the 12 communities did not find a single nontransitive trio, suggesting strongly hierarchical competition among our species. This discrepancy between theory and our findings may be caused by the underlying eco- logical interactions among competing spe- cies. In our communities, exploitation of the single externally supplied limiting nutrient and cross-feeding appeared to be the domi- nant ecological interactions determining the community structure (9, 28), whereas non- transitivity may emerge through interference competition (39) or through changes in spe- cies’ competitiveness across resources (40). Our experiments suggest that pairwise co- existence is not necessarily required for the stable assembly of multispecies communities. However, more complex assembly rules might still be found to predict and explain multispe- cies coexistence. Future empirical work with communities assembled under growing envi- ronmental complexity will be necessary to es- tablish how factors such as spatial structure, the number of supplied resources, the existence of higher-order interactions, and fluctuating conditions may influence the complexity of coexistence in multispecies communities. REFERENCES AND NOTES 1. G. E. Hutchinson, Am. Nat. 95, 137–145 (1961). 2. R. M. May, Nature 238, 413–414 (1972). 3. R. M. May, Stability and Complexity in Model Ecosystems (Princeton Univ. Press, 2019). 4. K. Z. Coyte, J. Schluter, K. R. Foster, Science 350, 663–666 (2015). J. D. Palmer, K. R. Foster, Science 376, 581–582 (2022). 5. 6. R. P. Rohr, S. Saavedra, J. Bascompte, Science 345, 1253497 (2014). 7. P. Chesson, Annu. Rev. Ecol. Syst. 31, 343–366 (2000). 8. T. A. Hoek et al., PLOS Biol. 14, e1002540 (2016). J. E. Goldford et al., Science 361, 469–474 (2018). 9. 10. E. Bairey, E. D. Kelsic, R. Kishony, Nat. Commun. 7, 12285 (2016). 11. J. M. Levine, J. Bascompte, P. B. Adler, S. Allesina, Nature 546, 56–64 (2017). 12. M. T. Angulo, A. Kelley, L. Montejano, C. Song, S. Saavedra, Nat. Ecol. Evol. 5, 1091–1101 (2021). 13. K. Z. Coyte, C. Rao, S. Rakoff-Nahoum, K. R. Foster, PLOS Biol. 19, e3001116 (2021). 14. J. Friedman, L. M. Higgins, J. Gore, Nat. Ecol. Evol. 1, 109 (2017). 15. T. Fukami, Annu. Rev. Ecol. Evol. Syst. 46, 1–23 (2015). 16. S. P. Hammarlund, T. Gedeon, R. P. Carlson, W. R. Harcombe, Nat. Commun. 12, 619 (2021). 17. S. Allesina, J. M. Levine, Proc. Natl. Acad. Sci. U.S.A. 108, 5638–5642 (2011). 18. R. A. Laird, B. S. Schamp, Am. Nat. 168, 182–193 (2006). 19. R. A. Laird, B. S. Schamp, Ecology 89, 237–247 (2008). 20. R. A. Laird, B. S. Schamp, J. Theor. Biol. 256, 90–95 (2009). 21. L. Gallien, N. E. Zimmermann, J. M. Levine, P. B. Adler, Ecol. Lett. 20, 791–800 (2017). 22. S. Soliveres et al., Ecol. Lett. 18, 790–798 (2015). 23. M. J. Liao, M. O. Din, L. Tsimring, J. Hasty, Science 365, 1045–1049 (2019). 24. B. Sinervo, C. M. Lively, Nature 380, 240–243 (1996). 25. I. Mizrahi, R. J. Wallace, S. Moraïs, Nat. Rev. Microbiol. 19, 553–566 (2021). 26. E. S. Wright, K. H. Vetsigian, Nat. Commun. 7, 11274 (2016). 27. L. M. Higgins, J. Friedman, H. Shen, J. Gore, bioRxiv 175737 [Preprint] (2017). 28. S. Estrela et al., Cell Syst. 13, 29–42.e7 (2021). 29. S. Estrela, Á. Sánchez, M. Rebolleda-Gómez, Front. Microbiol. 12, 657467 (2021). 30. J. Diaz-Colunga et al., Proc. Natl. Acad. Sci. U.S.A. 119, e2111261119 (2022). 31. S. Estrela, J. Diaz-Colunga, J. C. C. Vila, A. Sanchez-Gorostiaga, A. Sanchez, bioRxiv (2022), p. 2022.02.13.480281. 32. C. A. Serván, J. A. Capitán, J. Grilli, K. E. Morrison, S. Allesina, Nat. Ecol. Evol. 2, 1237–1242 (2018). 33. S. Estrela, A. Sanchez-Gorostiaga, J. C. Vila, A. Sanchez, eLife 10, e65948 (2021). 34. J. C. C. Vila, Y.-Y. Liu, A. Sanchez, ISME J. 14, 2505–2513 (2020). 35. J. Vila, vilacelestin/VilaLiuSanchez2020: Code and data for VilaLiuSanchez2020 (2020; https://zenodo.org/record/ 3817698). 36. T. N. Grainger, J. M. Levine, B. Gilbert, Trends Ecol. Evol. 34, 925–935 (2019). 37. E. C. Engel, J. F. Weltzin, Plant Ecol. 195, 77–85 (2008). 38. A. Conwill et al., Cell Host Microbe 30, 171–182.e7 (2022). 39. B. Kerr, M. A. Riley, M. W. Feldman, B. J. M. Bohannan, Nature 418, 171–174 (2002). 40. P. B. Rainey, M. Travisano, Nature 394, 69–72 (1998). Chang et al., Science 381, 343–348 (2023) 21 July 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 41. C.-Y. Chang, D. Bajic, J. Vila, E. Sylvie, A. Sanchez, Data for: Emergent coexistence in multispecies microbial communities,” Dryad (2023); https://doi.org/10.5061/dryad.bnzs7h4gb. 42. C.-Y. Chang, Code for: Emergent coexistence in multispecies microbial communities,” Zenodo (2023); https://zenodo.org/ record/8015230. ACKN OW LEDG MEN TS We thank the members of the Sanchez laboratory for helpful discussions. Funding: This work was partially funded by Young Investigator Award RGY0077/2016 from the Human Frontier Science Program to A.S. C.-Y.C. was supported by a Graduate Student Fellowship for Studying Abroad from the Ministry of Education, Taiwan. A.S. was partially supported by grant PID2021- 125478NA-100 funded by MCIN/AEI/10.13039/501100011033 and by “ERDF: A way of making Europe.” Author contributions: C.-Y.C., D.B., and A.S. conceived the idea and designed the study. C.-Y.C. performed the experiments. C.-Y.C. and D.B. analyzed the data. C.-Y.C., D.B., J.C.C.V., and S.E. provided reagents and data. C.-Y.C., D.B., J.C.C.V, S.E., and A.S. discussed the results and drafted the paper. C.-Y.C., D.B., and A.S. wrote the final version of the paper. Competing interests: The authors declare no competing interests. Data and materials availability: Processed and raw data can be accessed at Dryad (41). Code used to perform analyses can be accessed on Zenodo (42). Detailed methods for assembling the original communities used in this study are described in (9). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg0727 Materials and Methods Figs. S1 to S17 Tables S1 to S3 References (43–53) MDAR Reproducibility Checklist Submitted 2 December 2022; accepted 15 June 2023 10.1126/science.adg0727 Chang et al., Science 381, 343–348 (2023) 21 July 2023 6 of 6
10.1126_science.adg0844
RES EARCH ANTHROPOLOGY Cooperation across social borders in bonobos Liran Samuni1,2,3* and Martin Surbeck2,4* Cooperation beyond familial and group boundaries is core to the functioning of human societies, yet its evolution remains unclear. To address this, we examined grooming, coalition, and food-sharing patterns in bonobos (Pan paniscus), one of our closest living relatives whose rare out-group tolerance facilitates interaction opportunities between groups. We show that, as in humans, positive assortment supports bonobo cooperation across borders. Bonobo cooperative attitudes toward in- group members informed their cooperative relationships with out-groups, in particular, forming connections with out-group individuals who also exhibited high cooperation tendencies. Our findings show that cooperation between unrelated individuals across groups without immediate payoff is not exclusive to humans and suggest that such cooperation can emerge in the absence of social norms or strong cultural dispositions. M ore than any other species, humans cooperate across vast contexts, numbers, and social scales. We live in complex multilevel societies that promote the formation of strong cooperative rela- tionships not only with kin, allies, and friends but also with distant acquaintances and even strangers. The extent of human nonkin coop- eration is unmatched, with trade and sharing of commodities, knowledge, and skills (1–3) taking place not only within human residen- tial units (hereafter “groups”) but also across these units. Kin relations and repeated interaction op- portunities are important foundations of within- group cooperation across taxa (4, 5). However, these appear insufficient in explaining human large-scale cooperation (6). In humans, theoret- ical and empirical models consistently identify population structures, individual attributes, and cultural processes as critical components that promote cooperation (2, 7–13). These models predict that cooperation, such as the sharing of food resources, can emerge if the population structure permits clustering of similar individ- uals, for example, when cooperators co-reside or interact preferentially with one another (10, 11, 13). Tolerance and cooperation across groups are apparent in various animal taxa, including insects, birds, and mammals (14–17). However, the ability to cooperate with unrelated out- group partners with no immediate return is considered to be an exclusively human feature (18, 19). Human cooperation across residential groups facilitates resource and information transfer and the accumulation of knowledge, both of which support our long life spans and 1Cooperative Evolution Lab, German Primate Center, Göttingen, Germany. 2Department of Human Evolutionary Biology, Harvard University, Cambridge, MA, USA. 3School of Psychology and Neuroscience, University of St Andrews, St Andrews, UK. 4Department of Human Behavior, Ecology and Culture, Max Planck Institute for Evolutionary Anthropology, Leipzig, Germany. *Corresponding author. Email: lsamuni@dpz.eu (L.S.); msurbeck@fas.harvard.edu (M.S.) prolonged development and allow our species to thrive across the globe (2, 20, 21). But how distinctive is this human capacity? And what are the evolutionary foundations of human broad cooperation? Observations of the between-group rela- tions of bonobos (Pan paniscus), our closest living relatives together with chimpanzees, chal- lenge the notion that cooperation with no immediate return between distantly related individuals across groups is exclusive to humans. Within certain bonobo populations, individuals from distinct groups engage in a diverse range of interactions that span from aggression to cooperation, including grooming, forming alliances, and sharing high-value food resources (22–31). Despite prolonged and tolerant inter- actions between different groups, bonobos still maintain a clear in-group–out-group distinction (32). However, our understanding of bonobo cooperation across groups largely relies on des- criptive information, which lacks empirical ex- amination of the frequency, social structures, and underlying mechanisms that are involved. In this study, we characterized bonobo coop- eration across groups by observing within- and between-group interactions of 31 adults living in two social groups (Ekalakala: three adult males and eight adult females; and Kokoalongo: seven adult males and 13 adult females) in the Kokolopori Bonobo Reserve, Democratic Republic of the Congo (DRC). During a 2-year observation period, we documented 95 en- counters between the two groups, which lasted for around 20% of the total observation time. The durations of bonobo between-group en- counters varied substantially, from less than 1 hour to 14 consecutive days [12.5 ± 17.2 hours (mean ± SD)], which emphasizes that bonobo groups can associate in a nontransient manner, thereby facilitating opportunities for consistent cross-group exchange. Results Bonobo within- and between-group interac- tions consisted of a variety of cooperative acts, including 3744 grooming interactions, 592 coa- litions, 2920 cases of noncoalitionary aggressions, and 650 cases of food transfers (see supple- mentary materials for definitions). Between- group interactions represented 10% (N = 383, between 115 dyads) of all cases of grooming, 15% (N = 87, between 43 dyads) of all co- alitions, 14% (N = 402, between 126 dyads) of all noncoalitionary aggressions, and 6% (N = 41, between 28 dyads, including 16 donors and 15 recipients) of all food transfers observed. Although kin selection is a powerful driver of cooperation and between-group connections, interactions in this bonobo population are unlikely to be driven solely by genetic related- ness. Bonobos are a male philopatric species, and the only mother-offspring pairs in the two groups are four mothers and their sons who reside in the same group. Whereas female migration can produce close familial connec- tions across groups, permanent female immi- gration between the two groups has not been observed since the establishment of the research site in 2016, despite numerous (N = 22) immi- grant females arriving to or leaving from the two groups. Finally, analysis of 15-loci autosomal microsatellite genotypes revealed that only 6% of both within- and between-group dyads are second-degree relatives or higher (see supple- mentary materials). The bonobo cooperative acts vary in their susceptibility to cheating and delay of potential payoff. Grooming in bonobos is a frequent be- havior that necessitates little initial investment by actors and allows immediate return; it there- fore involves opportunities to both test partners’ willingness to cooperate and to receive imme- diate reward. Bonobo coalition formation re- quires joint action against a common opponent, which may provide benefits to all partners. Finally, food transfer (hereafter “sharing”) is an act that can incur an initial cost to actors (re- duced energetic or nutrient intake) and offers little guarantee of a future return, which re- sults in an uncertain payoff to actors. Increased defection opportunities in between-group food sharing relative to grooming are visible in our dataset. Across the 2 years of observation, food sharing was reciprocated only among 4 of the 28 between-group dyads that shared food. In comparison, grooming investment was highly reciprocated within and between groups (fig. S1), and >70% of between-group grooming in- teractions involved immediate return (within the same interaction). The variability of the risk that benefits will be reciprocated among these cooperative interactions provides a plat- form through which to explore the underlying structures of cooperation across borders in a nonhuman species. In human networks, population structures that allow for the positive assortment of coop- erators support the proliferation of coopera- tion by pooling the benefits among those who Samuni et al., Science 382, 805–809 (2023) 17 November 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E cooperate (9–11, 13). Cooperation assortment refers to the proclivity of individuals to selec- tively interact with others who have similar cooperative tendencies or traits and requires interaction strategies to be nonrandom. To investigate whether nonrandom assortment underlies bonobo cooperation across social borders, we first established that bonobos interacted nonrandomly and exhibited a pref- erence for specific partners both within and between groups. Specifically, using data per- mutations (see supplementary materials), we calculated the expected variance of bonobo within- and between-group grooming, coalition, and food-sharing interactions if interactions were randomly distributed among available partners and compared this variance against the “true” interaction variance (observed vari- ance). In accordance with assortativity predic- tions, the observed variance of all interaction types significantly exceeded what is expected by chance, both within and between groups (all P < 0.001; fig. S2). Partner selectivity in our bonobo population provides a basis upon which cooperation assortment can emerge. If bonobos selectively interact with partners who are more likely to cooperate, they can increase their chances for a positive net gain. To examine cooperation assortment, we operationalized bonobo coop- erative tendencies using interindividual var- iation in grooming, coalition formation, and food-sharing acts expressed within groups (fig. S3). We constructed a social network for each of the bonobo interactions and exam- ined whether individuals categorized as high cooperators within their groups are more likely to connect groups in the different net- works (Fig. 1). The social network analyses illustrate that bonobos who groom, form coalitions, or donate more food to in-group members are more likely to form the same kinds of connections with out-group members (Fig. 1 and fig. S4). By contrast, within-group cooperative tendencies within one form of cooperation do not appear to predict connections between groups in another form of cooperation (figs. S5 to S7). These pat- terns may emerge if bonobos have consistent cooperative tendencies within, rather than across, the different cooperation forms—whether inter- acting with in-group or out-group. It is unlikely that kin relations explain the observed coop- erative patterns between groups because none of the between-group dyads were parent- offspring and only 5, 12, and 11% of the dyads that groomed, formed a coalition, and shared food between the groups, respectively, were identified as at least second-degree relatives (see supplementary materials). Overall, the social networks suggest that bonobos who more frequently cooperate within their groups are (i) more likely to engage in the same behavior with out-group individuals relative to less-frequent within-group cooper- ators and (ii) engage with out-group members who are also frequent within-group coopera- tors (assortment of cooperators). To test these two characteristics of the bonobo networks, we used Bayesian Poisson regression models. First, we tested how individual within-group cooperative tendencies affected cooperative relationships with out-group members. Account- ing for interaction opportunities, we found that bonobos who showed higher within-group food-sharing or coalitionary tendencies were also more likely to form the same kind of connections with out-group individuals [food sharing: estimate = 0.74, 95% credible in- terval (CI95%) = 0.38 to 1.14, odds ratio 2.1 per 1–standard deviation (SD) increase, table S1; coalition: estimate = 1.11, CI95% = 0.78 to 1.47, odds ratio 3 per 1-SD increase, table S2]. We did not find the same pattern when using grooming as the cooperative behavior (table S3). Identical analyses conducted to examine the impact of the within-group cooperative tendencies on between-group interactions across behaviors supported the idea of consistent cooperative tendencies within, but not across, currencies Grooming Coalition Food sharing High within-group cooperator Medium within-group cooperator Low within-group cooperator Within-group Between-group Female Male Unrelated Kin Fig. 1. Cooperation assortment in bonobo social networks. Social networks indicate that high within-group cooperators tend to be located more centrally and form the same types of connections across groups. Within-group cooperative tendencies were operationalized for each cooperative interaction: (i) grooming network, by dividing the amount of time a bonobo groomed an in-group member (average individual grooming time: Ekalakala, 2268 ± 881 min; Kokoalongo, 704 ± 439 min) by the average grooming duration within their group; (ii) coalition network, by dividing the number of times a bonobo formed a coalition with an in-group member (average individual coalition: Ekalakala, 36 ± 27; Kokoalongo, 63 ± 53) by the average coalition times within their group; and (iii) food-sharing network, by dividing the number of times a bonobo donated food to an in-group member (average individual sharing: Ekalakala, 27 ± 18; Kokoalongo, 16 ± 15) by the average food-sharing times within their group. Node colors indicate high (66th percentile; red), medium (between the 33rd and 66th percentile; blue), and low (below the 33rd percentile; white) within-group cooperators. Node shape indicates males (square) or females (circles), and node size indicates the number of both incoming and outgoing connections possessed by an individual, with larger nodes signifying a greater number of connections. Edges represent connections with values equal to or greater than the population mean, with their colors indicating whether partners reside in the same (yellow) or different (gray) groups or whether partners are related (dashed line) or not (solid line). Only individuals that appeared in the data during both observation years (N = 27 out of the 31 individuals in the data) are depicted in the social network illustrations. The individuals that are not connected to the main network are three males from the Kokoalongo group (food-sharing network) and one male from Kokoalongo and three nulliparous females from Ekalakala (coalition network). Samuni et al., Science 382, 805–809 (2023) 17 November 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E (tables S1 to S3). Finally, we also found that bonobos who had higher sharing tendencies, but not higher grooming or coalition tenden- cies, within groups were less likely to engage in aggression with out-group members (estimate = −0.36, CI95% = −0.53 to −0.20; fig. S5 and table S4). Competition and aggression may jeopardize cooperative relationships, particularly with a delay between action and reward, as with food sharing. As such, the reduced likelihood of between-group aggressions in those with high food-sharing tendencies may offer a pathway to foster the maintenance of food-sharing rela- tionships across groups. Cooperation assortment can emerge if indi- viduals interact with those who are more able and willing to confer benefits (33). Accordingly, we would expect individuals who are higher cooperators within their groups to be more likely to cooperate with out-group members who are themselves higher cooperators. Therefore, in a second step, we tested whether the combined within-group cooperation tendencies of dyads (examining each cooperation form separately) predicted interactions between groups. Consis- tent with this prediction, cooperation patterns between groups were best explained by the sum of the within-group cooperative tendencies of partners (i.e., “cooperation score”). A 1-SD increase in the food sharing, coalition, or grooming cooperation scores increased between- group food sharing, coalition, or grooming odds by factors of 1.77, 2.63, and 1.35, respectively (Fig. 2, fig. S8, and table S5). Further, part- ners who groomed more frequently also had a higher relative sharing probability (esti- A Grooming B Food sharing Cooperation score Food sharing Coalition Female-Male Male-Male Cooperation score Grooming Coalition Female-Male Male-Male -4 -3 -2 -1 0 1 -4 -3 -2 -1 0 1 1.00 0.75 0.50 0.25 0.00 C e e r g e d - n i i g n m o o r G E e e r g e d - n i i g n m o o r G 1.00 0.75 0.50 0.25 0.00 Within-groups 0.75 0.50 0.25 Grooming out-degree 1.00 Between-groups 0.00 0.75 0.50 0.25 Grooming out-degree 1.00 D e e r g e d - n i g n i r a h S F e e r g e d - n i g n i r a h S 1.00 0.75 0.50 0.25 0.00 1.00 0.75 0.50 0.25 0.00 Within-groups 0.00 0.50 0.25 0.75 Sharing out-degree 1.00 Between-groups 0.0 0.1 0.2 0.3 Sharing out-degree Fig. 2. Positive assortment underlies bonobo cooperation across borders. (A and B) Estimates based on models that test whether the joint within-group cooperative tendencies of individuals (i.e., cooperation score) or interactions in other currencies explain their likelihood to groom (A) or share food (B) across groups. The cooperation score is calculated as the sum of the cooperative tendencies of a dyad based on either within-group grooming (A) or food donations (B). Shown are the model estimates (black diamonds) with 50% (yellow rectangles) and 95% (black lines) credible intervals derived from a Bayesian Poisson regression. (C to F) Generous bonobos are more attractive as interaction partners. The relationships between grooming [(C) and (E)] and food-sharing [(D) and (F)] in- and out-degree within [(C) and (D)] and between [(E) and (F)] groups are shown. The in-degree represents the number of partners from whom an individual received grooming or food, whereas the out-degree represents the number of partners that an individual groomed or donated food to. The in- and out-degrees are proportional to the potential number of partners that one can interact with within and between groups. Shown are the data points (N = 31 individuals; females are represented by circles and males by squares), model estimates (yellow, within-group; blue, between- group), and the 95% credible intervals (gray) derived from a Bayesian regression. mate = 0.33, CI95% = 0.05 to 0.61, odds ratio 1.39 per 1-SD increase), and vice versa (esti- mate = 0.35, CI95% = 0.25 to 0.46, odds ratio 1.42 per 1-SD increase). The sex combination of the partners had no obvious consistent impact on their cooperation patterns (Fig. 2, fig. S8, and table S5). The between-group connections of coopera- tors in the different networks confirm the pres- ence of cooperation assortment in bonobos and that reciprocity might be at play in sup- porting bonobo cooperative acts. Nonetheless, strong connections of high cooperators be- tween groups may also arise as a result of ran- dom processes. For example, in a population with random partner selection and varying interindividual sharing tendencies, food sharing is expected to accumulate between those indi- viduals who share more. The same is true for the other cooperative interactions. Such a process can generate stronger cross-group connections between high within-group cooperators by chance, without the need for preferential inter- actions. However, given that partner selection in Kokolopori bonobos is not random (see permutation procedure), the observed assort- ment of cooperators in the different networks is unlikely to be a mere artifact of random processes. Further, in the context of food shar- ing, that only four (14%) between-group dyads reciprocated food sharing further strengthens the idea that cooperation assortment is not a product of chance. Finally, cooperation can evolve when more- generous individuals are also more likely to obtain benefits, whether through reciprocity or because of their attractiveness as interac- tion partners. Subsequently, it is predicted that those who are more able or willing to benefit others will be more readily chosen by others as interaction partners. We therefore examined whether bonobos who groomed or donated food to more partners (“high out- degree”) also received grooming or food from more partners (“high in-degree”) by dividing the number of partners each individual groomed or donated food to or received food from by the total number of potential partners with whom one could have interacted. Following this pro- cedure, we generated values between 0 and 1, with 1 indicating that an individual interacted with all potential partners. We could not eval- uate the same question for coalition formation because there is no clear actor or receiver in this type of interaction. We found a strong relationship (Fig. 2) between the grooming and food sharing in- and out-degrees of bonobos both within groups [grooming: regression co- efficient (R) = 0.99; food sharing: R = 0.81] and between groups (grooming: R = 0.94; food sharing: R = 0.41). Overall, individuals that were more generous, by grooming more partners or donating food to more individuals, were also more likely to receive benefits from more Samuni et al., Science 382, 805–809 (2023) 17 November 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E individuals. The in- and out-degree grooming patterns were similarly strong within and be- tween groups, likely because of high grooming return in bonobos (fig. S1). In comparison, the relationship between food-sharing in- and out-degree was especially evident within groups, where repeated interaction opportunities, and hence reciprocity or knowledge about part- ners, are more certain (53% of within-group dyads shared food reciprocally versus 14% of between-group dyads). Discussion Bonobos, our closest living relative together with chimpanzees, maintain stable but variable cooperative relationships that transcend group boundaries. The bonobo population structure permits repeated interactions and partner se- lectivity, which are fundamental components for the emergence of cooperative relationships. Bonobo nonkin cooperation across groups in- cluded food sharing, a prosocial act with uncertain returns and a high defection proba- bility that is considered to be a key aspect of the human collaborative foraging niche. Food sharing in humans promotes cooperative rela- tionships that support our expensive life his- tories by offsetting the risk of food shortfalls (34). Although the relevance of nonkin food sharing to the survival and reproduction of bonobos is unknown, we find that, like humans, bonobo sharing networks rely on the positive assortment of cooperators. Nonetheless, the pro- cesses that support cooperation assortment in humans and bonobos likely differ. Our results suggest that various mechanisms, including reciprocal altruism (i.e., cooperative investments based on past return) and partner choice, may explain cooperation assortment between bonobo groups. Whereas reciprocal altruism requires repeated interactions be- tween the same partners and some form of bookkeeping, partner choice assumes stable interindividual differences in cooperative char- acteristics and that individuals have knowledge of the characteristics of potential partners to make effective choices (35). Both repeated interactions and knowledge about partners are nontrivial when interactions occur between members of different social groups. As such, the high group-membership fluidity, transient interactions, and large social network in human societies at present may limit the efficiency and explanatory power of these mechanisms. Instead, cultural processes and norm psychology are suggested to stabilize assortment and large- scale cooperation in humans (2, 13, 19, 20). In bonobos, the small social network relative to that of humans and the frequent associations between individuals from different groups (32) facilitate interactions and the accumu- lation of social knowledge beyond one’s own group (e.g., which out-group individuals are more able and willing to confer benefits than others). Such a social structure provides a basis upon which partner familiarity can enforce reciprocity beyond group boundaries and co- operation between groups can be sustained. Although social norms and culture are consid- ered the main cooperation mechanisms that sustain human large-scale cooperation, it is reasonable to assume that reciprocity may have played a more crucial role in our evolu- tionary past when living in smaller societies. Bonobo society offers a rare opportunity to study a social system in which individuals from different groups engage in resource and com- modity exchange. Our research builds upon previous experiments in captivity that show that bonobo food sharing with unrelated and unfamiliar individuals is not solely motivated by selfish interests or immediate rewards (30, 31). The convergence of evidence from both wild and captive studies suggests that the xenophilic tendencies of bonobos may be in- trinsic to the species as a whole. The tolerant and cooperative between-group relations of bonobos (22–26, 28, 32) stand in contrast to the ubiquitously hostile between- group relations and the strong in-group favoritism observed in their sister species, chimpanzees [Pan troglodytes (36–38)]. A leading hypothe- sis in Pan speciation posits that bonobos have experienced a selection against aggression [the self-domestication hypothesis (39, 40)], which has led to a reduction of in-group favoritism and an overall increase in prosocial tendencies toward others, whether in-group or out-group (41). Therefore, it is often assumed that bonobos are inclined to interact prosocially with every- one, both strangers and familiar individuals, and that it is their nondiscriminatory pro- social tendencies that permit out-group rela- tionships and cooperation. Although the marked prosocial tendencies that the Kokolopori bonobos exhibit toward out-groups support the self- domestication hypothesis, our findings addi- tionally suggest that bonobo prosocial tendencies are discriminatory. Instead of generally high prosocial tendencies, we propose that it is the strategic social ties that bonobos form with those who are more able and likely to confer benefits to others that stabilizes bonobo cooperation across borders. Bonobos are not the only nonhuman species that exhibit cooperative relationships between nonkin across groups (14–17). For example, in bottlenose dolphins, males form “third-order alliances” with unrelated out-group males that allow them to successfully compete over fe- males (16). Third-order social alliances in dol- phins are akin to the bonobo coalitions that are formed between males and females of different groups. In both species, these inter- actions confer improved access to contested resources to all allies and/or increased social status, which is therefore better categorized as mutualism. However, bonobo cooperation across borders also includes a behavior that cannot be explained by mutualism, thereby incorporating cooperation aspects that are considered exclusive to humans, such as the ability to act prosocially toward unrelated out-group members with no guarantee of a return. Owing to habituation status and data col- lection constraints, our study only included two out of at least four bonobo groups that are ob- served to regularly interact within the Kokolopori population (32), resulting in an investigation of a relatively small social network (31 indi- viduals). Although these bonobos maintain a wider social network than that represented here, the overall size of this bonobo social network is limited and residential mobility is considerably reduced compared with humans (32), making it challenging to examine coopera- tion mechanisms of bonobos and humans on an equal footing. Further, unlike bonobos, cooperation across groups in humans extends beyond pair-wise interactions to also include large-scale cooperation (1–3). Nevertheless, our bonobo results suggest that the higher-order social connections and cooperative pair-wise relationships that humans form across groups may have a different evolutionary history than is often assumed. Theories in human evolution posit that pair bonding, exogamy, and the ability to recognize maternal and paternal kin and affines (i.e., relatives by marriage) are necessary compo- nents of between-group bonds and cooperation (19, 42). Here, bonobos offer an alternative scenario, in which cooperative ties between groups are formed in the absence of exogamy or strong bilateral kin recognition. It is there- fore plausible that an ancestral state of human between-group, pair-wise cooperation is that of a bonobo-like social system, in which toler- ance toward out-groups facilitates the emergence of cooperation in the absence of high degrees of genetic relatedness. Consequently, bonobos offer a key comparative model to human social systems and a rare opportunity to reconstruct the ancestral conditions of human large-scale cooperation. REFERENCES AND NOTES 1. R. Boyd, P. J. Richerson, Evol. Anthropol. 31, 175–198 2. (2022). J. Henrich, M. Muthukrishna, Annu. Rev. Psychol. 72, 207–240 (2021). 3. M. B. Brewer, Am. Psychol. 62, 726–738 (2007). 4. R. Axelrod, W. D. Hamilton, Science 211, 1390–1396 (1981). 5. T. Clutton-Brock, Nature 462, 51–57 (2009). 6. R. Boyd, P. J. Richerson, J. Theor. Biol. 132, 337–356 7. (1988). I. Eshel, L. L. Cavalli-Sforza, Proc. Natl. Acad. Sci. U.S.A. 79, 1331–1335 (1982). 8. S. Bowles, Science 314, 1569–1572 (2006). 9. M. Dyble et al., Curr. Biol. 26, 2017–2021 (2016). 10. C. L. Apicella, F. W. Marlowe, J. H. Fowler, N. A. Christakis, Nature 481, 497–501 (2012). 11. A. Romano, D. Balliet, T. Yamagishi, J. H. Liu, Proc. Natl. Acad. Sci. U.S.A. 114, 12702–12707 (2017). Samuni et al., Science 382, 805–809 (2023) 17 November 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E 12. H. Ohtsuki, C. Hauert, E. Lieberman, M. A. Nowak, Nature 441, 502–505 (2006). 31. J. Tan, D. Ariely, B. Hare, Sci. Rep. 7, 14733 (2017). 32. L. Samuni, K. E. Langergraber, M. H. Surbeck, Proc. Natl. Acad. 13. C. Handley, S. Mathew, Nat. Commun. 11, 702 (2020). 14. E. J. H. Robinson, J. L. Barker, Biol. Lett. 13, 20160793 Sci. U.S.A. 119, e2201122119 (2022). 33. K. M. Smith, C. L. Apicella, Evol. Hum. Behav. 41, 354–366 (2017). 15. C. C. Grueter et al., Trends Ecol. Evol. 35, 834–847 (2020). 16. R. C. Connor, M. Krützen, S. J. Allen, W. B. Sherwin, S. L. King, Proc. Natl. Acad. Sci. U.S.A. 119, e2121723119 (2022). I. Krams, T. Krama, K. Igaune, R. Mänd, Behav. Ecol. Sociobiol. 62, 599–605 (2008). 17. 18. E. Fehr, U. Fischbacher, Nature 425, 785–791 (2003). 19. C. L. Apicella, J. B. Silk, Curr. Biol. 29, R447–R450 (2020). 34. H. S. Kaplan, P. L. Hooper, M. Gurven, Philos. Trans. R. Soc. Lond. Ser. B 364, 3289–3299 (2009). 35. R. McElreath, R. Boyd, Mathematical Models of Social Evolution: A Guide for the Perplexed (Univ. Chicago Press, 2007). 36. R. W. Wrangham, Am. J. Phys. Anthropol. 110, 1–30 (1999). 37. M. L. Wilson et al., Nature 513, 414–417 (2014). 38. L. Samuni, C. Crockford, R. M. Wittig, Nat. Commun. 12, 539 (2019). (2021). 20. R. Boyd, P. J. Richerson, Philos. Trans. R. Soc. Lond. Ser. B 39. B. Hare, V. Wobber, R. Wrangham, Anim. Behav. 83, 573–585 364, 3281–3288 (2009). 21. A. B. Migliano et al., Sci. Adv. 6, eaax5913 (2020). 22. L. Cheng et al., Horm. Behav. 128, 104914 (2021). 23. T. Furuichi, Int. J. Primatol. 41, 203–223 (2020). 24. G. Itani, Afr. Study Monogr. 11, 153–186 (1990). 25. T. Sakamaki, H. Ryu, K. Toda, N. Tokuyama, T. Furuichi, Int. J. Primatol. 39, 685–704 (2018). 26. N. Tokuyama, T. Sakamaki, T. Furuichi, Am. J. Phys. Anthropol. 170, 535–550 (2019). 27. L. Cheng, L. Samuni, S. Lucchesi, T. Deschner, M. Surbeck, Anim. Behav. 187, 319–330 (2022). 28. B. Fruth, G. Hohmann, Hum. Nat. 29, 91–103 (2018). 29. L. Samuni, F. Wegdell, M. Surbeck, eLife 9, e59191 (2020). 30. J. Tan, B. Hare, PLOS ONE 8, e51922 (2013). (2012). 40. R. W. Wrangham, Proc. Natl. Acad. Sci. U.S.A. 115, 245–253 (2018). 41. J. Tan, B. Hare, in Bonobos, B. Hare, S. Yamamoto, Eds. (Oxford Univ. Press, 2017), pp. 140–154. 42. B. Chapais, Evol. Anthropol. 22, 52–65 (2013). 43. L. Samuni, LiranSamuni/Bonobo-cooperation-across-borders: Cooperation across borders in bonobos. Zenodo (2023); https://doi.org/10.5281/zenodo.8156761. ACKN OWLED GMEN TS We thank the Institut Congolais pour la Conservations de la Nature (ICCN) and the Ministry of Scientific Research and Technology in the DRC for their support and permission to work in the Kokolopori Bonobo Reserve, DRC. We also thank the Bonobo Conservation Initiative and Vie Sauvage, especially S. Coxe, A. Lotana Lokasola, A. Menante, and staff members of the Kokolopori Bonobo Research Project for supporting our work. We thank L. Vigilant and V. Städele for conducting the genetic analysis and J. Henrich, C. Curtin, and E. Wessling for helpful discussions and comments. Funding: This work was funded by Harvard University (M.S.), the Max Planck Society (M.S., L.S.), the British Academy (L.S.), and Deutsche Forschungsgemeinschaft (L.S.). Author contributions: Conceptualization: L.S., M.S.; Methodology: L.S., M.S.; Investigation: L.S.; Funding acquisition: L.S., M.S.; Project administration: M.S.; Writing – original draft: L.S.; Writing – review and editing: L.S., M.S. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All processed data and code used in the analyses are available on Zenodo (43). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg0844 Materials and Methods Figs. S1 to S9 Tables S1 to S5 References (44–55) MDAR Reproducibility Checklist Submitted 2 December 2022; accepted 29 September 2023 10.1126/science.adg0844 Samuni et al., Science 382, 805–809 (2023) 17 November 2023 5 of 5
10.1126_science.adg3812
150-mm-thick lithium niobate optical reso- nator placed inside a superconducting alu- minum microwave cavity at a temperature of 7 mK, described in detail in (24). The micro- wave mode âe is electro-optically coupled to the colocalized optical whispering gallery modes at wo/2p ≈ 193.46 THz through the Pockels effect (Fig. 1A). The microwave reso- nance frequency we/2p is tuned to the optical free spectral range (FSR) of 8.799 GHz to re- alize a triply resonant system with interaction Hamiltonian: ^H int ¼ ħg0^ap^ae †^ao † þ h:c: ð1Þ p ^a† e (cid:3) ffiffiffiffiffi (cid:2)np † þ ^ae^ao ^ao where g0 is the vacuum electro-optical coupling rate and âp (âo) is the annihilation operator of the optical pump (Stokes) mode (25–28). Herein, we have ignored the interaction with the sup- pressed optical anti-Stokes mode ât (Fig. 1B) (29). In this sideband suppressed situation, two- mode squeezing interaction was achieved with a strong resonant optical pump tone, yield- ing the simple effective Hamiltonian, ^H eff ¼ (cid:4) †^api is , where (cid:2)np ¼ h^ap ħg0 the mean intracavity photon number of the op- tical pump mode. Entanglement between the out-propagating microwave and the optical field was generated through spontaneous parame- tric down-conversion (SPDC) below the para- metric instability threshold (C < 1). Here, C ¼ 4(cid:2)npg2 Þ is the cooperativity with the vac- 0 uum coupling rate g0/2p ≈ 37 Hz and the total loss rates of the microwave and optical Stokes modes ke/2p ≈ 11 MHz and ko/2p ≈ 28 MHz. The required ultralow-noise operation for en- tanglement generation was achieved in the pulsed regime because of the slow heating of this millimeter-sized device despite the required high-power optical pump (24). ð = keko Establishing nonclassical correlations We characterized the microwave and optical output fields using continuous variables, i.e., RES EARCH QUANTUM OPTICS Entangling microwaves with light R. Sahu1*†, L. Qiu1*†, W. Hease1‡, G. Arnold1, Y. Minoguchi2, P. Rabl2,3,4,5, J. M. Fink1* Quantum entanglement is a key resource in currently developed quantum technologies. Sharing this fragile property between superconducting microwave circuits and optical or atomic systems would enable new functionalities, but this has been hindered by an energy scale mismatch of >104 and the resulting mutually imposed loss and noise. In this work, we created and verified entanglement between microwave and optical fields in a millikelvin environment. Using an optically pulsed superconducting electro-optical device, we show entanglement between propagating microwave and optical fields in the continuous variable domain. This achievement not only paves the way for entanglement between superconducting circuits and telecom wavelength light, but also has wide-ranging implications for hybrid quantum networks in the context of modularization, scaling, sensing, and cross-platform verification. T he ability to manipulate and measure quantum mechanical properties such as quantum superpositions and entangle- ment in a variety of physical systems serves as the basis for the development of quantum technologies, for which the dem- onstration of quantum advantage with tens of superconducting qubits (1), an ultracoher- ent quantum memory with nuclear spins (2), and distributed quantum entanglement over tens of kilometers using optical photons (3) represents just a few of the highlights that have already been achieved. Combining these techniques (4–6) will enable the realization of general-purpose quantum networks in which remote quantum nodes capable of storing and processing quantum information seamlessly communicate with each other by distribut- ing entanglement over optical channels (7). Aspects of this approach have already been adopted to connect and entangle various quan- tum platforms remotely. These involve single atoms, ions, atomic ensembles, quantum dots, rare-earth ions, and nitrogen-vacancy centers (8). However, such long-distance quantum connectivity is considerably more difficult to achieve with microwave-based platforms such as semiconductor spin qubits (9) or local cryo- genic networks of superconducting circuits (10, 11), for which no natural interface to prop- agating optical photons is available. To overcome this limitation, efforts are cur- rently focused on the development of coher- ent quantum transducers between microwave and optical photons (12–17). Direct noiseless 1Institute of Science and Technology Austria, am Campus 1, 3400 Klosterneuburg, Austria. 2Vienna Center for Quantum Science and Technology, Atominstitut, TU Wien, 1040 Vienna, Austria. 3Walther-Meißner-Institut, Bayerische Akademie der Wissenschaften, 85748 Garching, Germany. 4Technische Universität München, TUM School of Natural Sciences, 85748 Garching, Germany. 5Munich Center for Quantum Science and Technology (MCQST), 80799 Munich, Germany. *Corresponding author. Email: rsahu@ist.ac.at (R.S.); liu.qiu@ ist.ac.at (L.Q.); jfink@ist.ac.at (J.M.F.) †These authors contributed equally to this work. ‡Present address: Quandela SAS, 91120 Palaiseau, France. conversion of a quantum state typically relies on a beam-splitter process in which a strong driving field mediates the conversion be- tween weak microwave and optical signals. However, a deterministic channel based on direct transduction has exceptionally strin- gent requirements on conversion efficiency and added classical noise that are still out of reach. Alternatively, quantum states can be teleported over long distances (18–20) by first establishing an entangled state between microwave and optical photons. Because they are assisted by an error-free classical signal (21, 22), such teleportation-based commu- nication channels are more tolerant to noise and losses, and their channel capacity is never zero as long as a finite amount of entangle- ment can be shared between the communi- cating nodes (22). Theoretical outline and experimental setup We generated and verified entanglement be- tween microwave and optical fields in the continuous-variable domain (23) using an ultralow-noise cavity electro-optical device (21). Our device consists of a 5-mm-diameter, Fig. 1. Physical and conceptual mode configuration. (A) Simulated microwave (left) and optical (right) mode distribution with azimuthal number me = 1 and mo = 17 (for illustration, experimentally mo ≈ 20,000). Phase matching is fulfilled from the condition mo = mp – me, and entanglement is generated and verified between the out-propagating microwave field âe,out and the optical Stokes field âo,out. (B) Sketch of the density of states of the relevant modes. Under the condition wp – wo = we, the strong pump tone in âp produces entangled pairs of microwave and optical photons in âe and âo through spontaneous parametric down-conversion. Frequency up- conversion is suppressed through hybridization of the anti-Stokes mode ât with an auxiliary mode. Sahu et al., Science 380, 718–721 (2023) 19 May 2023 1 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Measurement sequence and noise powers. (A) Schematic pulse sequence of a single measurement. The optical pulse 1 is applied at wp and amplifies the vacuum (and any thermal noise) in the two modes âe and âo, thus generating the SPDC signals. One micro-second later, a second optical pump with ~10 times lower power is applied, together with a coherent microwave pulse at we. The microwave photons stimulate the optical pump to down-convert, which generates a coherent pulse in the âo mode that is used to extract slow local oscillator phase drifts. (B and C) Measured output power from the âe and âo modes in units of photons per second in a 1-Hz bandwidth and averaged over a million experiments. The SPDC signals are shown in the insets, with the dashed gray lines indicating the calibrated detection noise floor Nj,add + 0.5. (D) Corresponding microwave output power spectral density versus Dwe = w – we centered on resonance right before the entanglement pulse, during the pulse, and right after the pulse, as indicated in (A). Yellow and green dashed lines are fits to a Lorentzian function that yields the microwave bath occupancies before and after the entangling pulse. Error bars represent the 2s statistical SE, and the shaded regions represent the 95% confidence interval of the fit. (E) Corresponding optical output power spectral density versus Dwo = wo – w during and after the entanglement pulse, both normalized to the measured noise floor before the pulse. The in-pulse noise spectra in (D) and (E) are fit jointly with theory, which yields C = 0.18 ± 0.01 and (cid:2)ne;int ¼ 0:07 T 0:03. ð the dimensionless quadrature pairs Xj and Pj (where j = e,o for microwave and optics), which satisfy the canonical commutation rela- tions [Xj, Pj] = i. A pair of Einstein-Podolsky- Rosen–type operators, Xþ ¼ 1ffiffi Þ and p Xe þ Xo ð 2 P(cid:2) ¼ 1ffiffi Þ, were then constructed (30). p Pe (cid:2) Po 2 The microwave and optical output fields are entangled if the variance of the joint op- erators is reduced below the vacuum level, i.e., D(cid:2) (cid:2)i < 1. This is common- ly referred to as the Duan-Simon criterion (31, 32). Here, the entanglement was estab- lished deterministically between the quad- ratures of two propagating bosonic modes and verified through the measured quad- rature variances based on all collected data (33). This is in contrast to probabilistic en- tanglement in the single photon basis, which is established with finite probability only after þ þ P2 ¼ hX 2 EPR successful heralding from an auxiliary mea- surement (34). For entanglement generation, we used a 250-ns-long optical pump pulse (≈154 mW, C ≈ 0.18, (cid:2)np ≈ 1:0 (cid:3) 1010 ) at a 2-Hz repeti- tion rate (pulse 1 in Fig. 2A). The output optical signal was filtered through a Fabry- Perot cavity to reject the strong pump. The microwave output field was amplified by a high-electron-mobility transistor amplifier. Both outputs were down-converted to an in- termediate frequency of 40 MHz with two local oscillators, and the four quadratures were extracted from heterodyne detection. Long-term phase stability between the two local oscillators was achieved by extraction of the relative phase drift by means of a sec- ond phase-alignment pump pulse that was applied 1 ms after each entanglement pulse, together with a coherent resonant microwave pulse (Fig. 2A). This generated a high signal- to-noise coherent optical signal through stim- ulated parametric down-conversion and allowed for phase alignment in each indi- vidual measurement (29). Figure 2, B and C, shows the time-domain power over 1 million averages for the on- resonance microwave and optics signal with a 40-MHz measurement bandwidth. The two insets show the microwave (optical) signal from spontaneous parametric down-conversion (SPDC) from pulse 1 (compare Fig. 2A) with an emission bandwidth of ≈10 MHz. The larger signals during the second pump pulse are the reflected microwave pulse and the generated optical tone, and these were used for phase alignment. The off-resonance raw power measurements were rescaled to the Sahu et al., Science 380, 718–721 (2023) 19 May 2023 2 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Characterization of the two-mode squeezed state. (A) Measured covariance matrix Vij in its standard form plotted for Dwj = 0 based on 925,000 measurements. (B) Corresponding Wigner function marginals of different output quadrature pairs compared with vacuum. The contours in blue (gray) represent the 1/e fall-off from the maximum for the measured state (vacuum). The middle two plots show two-mode squeezing below the vacuum level in the diagonal and off-diagonal directions. (C) Top, measured average microwave output noise Þ=2 (purple), average optical output noise (cid:2)V11 ¼ V11 þ V22 ð (cid:2)V13 ¼ V13 (cid:2) V24 ð Þ=2 (green), and average correlations (cid:2)V33 ¼ V33 þ V44 ð (yellow) as a function of measurement detunings. Solid lines represent the joint theory fit, and dashed lines are individual Lorentzian fits to serve as a guide to the eye. Middle bottom, two-mode squeezing in red (antisqueezing in blue) (cid:2)V13. The darker-colored calculated from the top panels as DT error bars represent the 2s statistical error, and the outer (faint) 2s error bars also include the systematic error in calibrating the added noise of the measurement setup. ¼ (cid:2)V11 þ (cid:2)V33 T 2 Þ=2 EPR detection noise floor Nj,add + 0.5, with the added noise Ne,add = 13.1 ± 0.4 (2s errors throughout) due to loss and amplifier noise and No,add = 5.5 ± 0.2 due to optical losses, which were carefully determined using noise thermometry of a temperature-controlled 50 W load and four-port calibration, respectively (29). Because the emission bandwidth was smaller than the measurement bandwidth, the time domain measurement does not reveal the full SPDC amplitude. We continued the analysis in the frequency domain by calculating the Fourier transform of each measurement for three separate time intervals: before (2 ms), during (200 ns), and immediately after (500 ns) the entangling pump pulse. Figure 2D shows the resulting average microwave noise spectra for all three time in- tervals with corresponding fit curves (dashed lines) and theory (solid line). The spectra were in situ calibrated using the before-pulse off- resonance (waveguide noise) noise floor (29). Using independent measurements, we deter- mined this noise floor (cid:2)ne;wg ¼ 0:001 T 0:002 at the low average pump power of ≈0.12 mW used in this experiment (29). Before and after the pump pulse, the intrinsic microwave bath occu- pancies were fitted to be (cid:2)ne;int ¼ 0:03 T 0:01 and 0.09 ± 0.03, respectively, above the vacuum level. Similarly, Fig. 2E shows the obtained av- erage optical noise spectra during and after the pump pulse, calibrated through the shot noise level in the heterodyne measurement before the pulse. As expected, the optical noise level after the pulse returned back to the shot noise level. During the pump pulse, approximately Lorentzian-shaped microwave and optical power spectra were generated through the SPDC process. A joint fit of the microwave and op- tical power spectral density during the pulse was performed using a five-mode theoretical model that includes the effects of measure- ment bandwidth. In this model, the in-pulse microwave bath occupancy (cid:2)ne;int ¼ 0:07 T 0:03 and the cooperativity C = 0.18 ± 0.01 are the only free fit parameters (29). The narrowed microwave linewidth ke,eff/2p = 9.8 ± 1.8 MHz (taken from a Lorentzian fit) agrees with co- herent electro-optical back-action (28). From the extracted numbers, we conclude that dur- ing the entanglement pulse, the quantum noise dominates the intrinsic microwave thermal noise, a prerequisite for microwave-optics en- tanglement generation. The quadratures Xj and Pj were extracted as a function of frequency, i.e., at Dwj = ±(w – wj) around the resonances due to energy conser- vation in the SPDC process (29, 35) during the pump pulse. At each frequency, the bi- partite Gaussian state of the propagating out- put fields was fully characterized by the 4 × 4 covariance matrix (CM) Vij = hduiduj + dujduii/2, where dui = ui – huii and u ∈ {Xe, Pe, Xo, Po} (29). The CM corresponds to the quantum state of the propagating fields in the coaxial line and the coupling prism at the device out- put, i.e., before setup losses or amplification incur. The diagonal elements in V correspond to the individual output field quadrature var- iances in dimensionless units, which can be obtained by subtracting the added detection Sahu et al., Science 380, 718–721 (2023) 19 May 2023 3 of 4 RES EARCH | R E S E A R C H A R T I C L E noise offsets from the measured variances, i.e., Vii(Dw) = Vii,meas(Dw) – Ni,add. The obtained CM from the data in Fig. 2 at Dw = 0 is shown in Fig. 3A in its standard form (31, 32). The nonzero off-diagonal elements indicate strong correlations between micro- wave and optical quadratures. To verify the quantum correlation, the two-mode squeezed quadratures are more intuitively visualized in terms of the quasiprobability Wigner function: (cid:5) exp (cid:2) 1 p 2 p2 uV (cid:2)1uT ffiffiffiffiffiffiffiffiffiffiffiffiffiffi det Vð Þ W uð Þ ¼ ð2Þ (cid:6) where u = (Xe, Pe, Xo, Po). Different marginals of this Wigner function are shown in Fig. 3B. The marginals from the same quadratures (Xe, Pe) and (Xo, Po) show uncorrelated ther- mal noise above the vacuum noise (gray circle) from SPDC. The cross-quadrature marginals (Xe, Xo) and (Pe, Po) show two-mode squeezing in the diagonal and off-diagonal directions be- low the vacuum level. The amount of squeez- ing was slightly different between the two because of the statistical uncertainty in the measured CM. ð ð Figure 3C shows the amount of two-mode squeezing between microwave and optical out- put fields. The averaged microwave quadra- ture variance (purple dots) (cid:2)V 11 ¼ V11 þ V22 Þ=2 and the averaged optics quadrature variance (green dots) (cid:2)V 33 ¼ V33 þ V44 Þ=2 are shown in the top panel, along with the prediction from the five-mode theory (solid line) and a simple fit to a Lorentzian function (dashed line), which show perfect agreement. Measured microwave-optical correlations (yellow dots) (cid:2)V 13 ¼ V13 (cid:2) V24 Þ=2 and the Lorentzian fit ð (dashed line) lie slightly below the theoret- ical prediction (solid line), which we attrib- ute to remaining imperfections in the phase stability (29). EPR EPR The bottom two panels of Fig. 3C show the squeezed and antisqueezed joint quadra- ¼ (cid:2)V 11 þ (cid:2)V 33 ∓ 2 (cid:2)V 13 (red ture variances D∓ and blue, respectively). We observed two- mode squeezing below the vacuum level, i.e., D(cid:2) < 1, with a bandwidth close to the ef- fective microwave linewidth. The maximal two-mode squeezing of 0:72þ0:31 (cid:2)0:25 dB is ob- served on resonance where D(cid:2) ¼ 0:85þ0:05 (cid:2)0:06 EPR (2s, 95% confidence) obtained from ~1 mil- lion pulses with (cid:2)V 11 ¼ 0:93, (cid:2)V 33 ¼ 0:84, and (cid:2)V 13 ¼ 0:46. The reported 2s error on D(cid:2) EPR takes both statistical and systematic errors into account. Thus, the value of D(cid:2) EPR beats the classical limit ( D(cid:2) ¼ 1 ) by >5s (29). The measured two-mode squeezing signifies an itinerant microwave-optical entangled state with a logarithmic negativity of EN = 0.17. The supplementary materials contain additional data for longer pulses and varying optical pump power, which corroborate the presented re- sults and findings (29). EPR Conclusions The demonstration of deterministic quantum entanglement between propagating microwave and optical photons establishes a nonclassical communication channel between circuit quan- tum electrodynamics and quantum photonics. The achieved entanglement generation rate of ≈0.11 ebits/200-ns-long pulse (29) is in prac- tice limited by the slow pulse repetition rate. We expect orders-of-magnitude higher rates with improved thermalization, higher micro- wave and optical quality factors, and electro- optic coupling enhancements that reduce the required pump power and the associated thermal load. Coupling efficiency improve- ments will allow for higher levels of two-mode squeezing (29) and facilitate deterministic entanglement distribution schemes to qubits (29, 36), teleportation-based state transfer (21, 22, 29), and quantum-enhanced remote detection (37). This device and state prepa- ration scheme can also be used directly for probabilistic heralding assisted protocols (7, 38, 39) when the cooperativity is some- what reduced. This is the most promising way forward to mitigating optical setup losses and extending the entanglement to room- temperature fiber optics. Being fully com- patible with superconducting qubits in a millikelvin environment, such a device will facilitate the integration of remote supercon- ducting quantum processors into a single co- herent optical quantum network. This is relevant not only for modularization and scaling (40, 41) but also for efficient cross- platform verification of classically intractable quantum processor results (42). RE FERENCES AND NOTES F. Arute et al., Nature 574, 505–510 (2019). 1. 2. M. Zhong et al., Nature 517, 177–180 (2015). 3. Y. Yu et al., Nature 578, 240–245 (2020). 4. Z.-L. Xiang, S. Ashhab, J. Q. You, F. Nori, Rev. Mod. Phys. 85, 623–653 (2013). 5. G. Kurizki et al., Proc. Natl. Acad. Sci. U.S.A. 112, 3866–3873 (2015). 6. A. A. Clerk, K. W. Lehnert, P. Bertet, J. R. Petta, Y. Nakamura, Nat. Phys. 16, 257–267 (2020). 7. L.-M. Duan, M. D. Lukin, J. I. Cirac, P. Zoller, Nature 414, 413–418 (2001). 8. S.-H. Wei et al., Laser Photonics Rev. 16, 2100219 (2022). 9. G. Burkard, T. D. Ladd, J. M. Nichol, A. Pan, J. R. Petta, arXiv:2112.08863 [cond-mat.mes-hall] (2021). 10. P. Magnard et al., Phys. Rev. Lett. 125, 260502 (2020). 11. Y. Zhong et al., Nature 590, 571–575 (2021). 12. X. Han, W. Fu, C.-L. Zou, L. Jiang, H. X. Tang, Optica 8, 1050 (2021). 13. M. Mirhosseini, A. Sipahigil, M. Kalaee, O. Painter, Nature 588, 599–603 (2020). 14. R. Sahu et al., Nat. Commun. 13, 1276 (2022). 15. B. M. Brubaker et al., Phys. Rev. X 12, 021062 (2022). 16. A. Kumar et al., Nature 615, 614–619 (2023). 17. M. J. Weaver et al., arXiv:2210.15702 [quant-ph] (2022). 19. K. G. Fedorov et al., Sci. Adv. 7, eabk0891 (2021). 20. S. L. N. Hermans et al., Nature 605, 663–668 (2022). 21. A. Rueda, W. Hease, S. Barzanjeh, J. M. Fink, npj Quantum Inf. 5, 108 (2019). 22. J. Wu, C. Cui, L. Fan, Q. Zhuang, Phys. Rev. Appl. 16, 064044 (2021). 23. U. L. Andersen, T. Gehring, C. Marquardt, G. Leuchs, Phys. Scr. 91, 053001 (2016). 24. W. Hease et al., PRX Quantum 1, 020315 (2020). 25. M. Tsang, Phys. Rev. A 81, 063837 (2010). 26. V. S. Ilchenko, A. A. Savchenkov, A. B. Matsko, L. Maleki, J. Opt. Soc. Am. B 20, 333 (2003). 27. L. Fan et al., Sci. Adv. 4, eaar4994 (2018). 28. L. Qiu, R. Sahu, W. Hease, G. Arnold, J. M. Fink, arXiv:2210.12443 [quant-ph] (2022). 29. See the supplementary materials. 30. A. Einstein, B. Podolsky, N. Rosen, Phys. Rev. 47, 777–780 (1935). 31. L.-M. Duan, G. Giedke, J. I. Cirac, P. Zoller, Phys. Rev. Lett. 84, 2722–2725 (2000). 32. R. Simon, Phys. Rev. Lett. 84, 2726–2729 (2000). 33. Z. Y. Ou, S. F. Pereira, H. J. Kimble, K. C. Peng, Phys. Rev. Lett. 68, 3663–3666 (1992). 34. H. Bernien et al., Nature 497, 86–90 (2013). 35. S. Zippilli, G. D. Giuseppe, D. Vitali, New J. Phys. 17, 043025 (2015). 36. J. Agustí, Y. Minoguchi, J. M. Fink, P. Rabl, Phys. Rev. A 105, 062454 (2022). 37. S. Barzanjeh et al., Phys. Rev. Lett. 114, 080503 (2015). 38. C. Zhong et al., Phys. Rev. Lett. 124, 010511 (2020). 39. S. Krastanov et al., Phys. Rev. Lett. 127, 040503 (2021). 40. S. Bravyi, O. Dial, J. M. Gambetta, D. Gil, Z. Nazario, J. Appl. Phys. 132, 160902 (2022). 41. J. Ang et al., arXiv:2212.06167 [quant-ph] (2022). 42. J. Knörzer, D. Malz, J. I. Cirac, arXiv:2212.07789 [quant-ph] (2022). 43. R. Sahu, Data for: Entangling microwaves with light, Zenodo; https://doi.org/10.5281/zenodo.7789418. AC KNOWLED GME NTS L.Q. thanks J. Li and D. Vitali for fruitful discussions. Funding: This work was supported by the European Research Council (grant no. 758053, ERC StG QUNNECT) and the European Union’s Horizon 2020 Research and Innovation Program (grant no. 899354, FETopen SuperQuLAN). L.Q. acknowledges generous support from the ISTFELLOW program. W.H. is the recipient of an ISTplus postdoctoral fellowship with funding from the European Union’s Horizon 2020 Research and Innovation Program (Marie Sklodowska-Curie grant no. 754411). G.A. is the recipient of a DOC fellowship of the Austrian Academy of Sciences at IST Austria. J.M.F. acknowledges support from the Austrian Science Fund (FWF) through BeyondC (grant no. F7105) and the European Union’s Horizon 2020 Research and Innovation Program (grant no. 862644, FETopen QUARTET). Author contributions: R.S., W.H., L.Q., and G.A. worked on the experimental setup. R.S. and L.Q. performed the measurements. L.Q. and R.S. analyzed the data. L.Q. developed the theory with contributions from R.S., Y.M., and P.R. R.S. and L.Q. wrote the manuscript with contributions from all authors. J.M.F. supervised the project. Competing interests: The authors declare no competing interests. Data and materials availability: All data are available in the manuscript or the supplementary materials or have been deposited on Zenodo (43). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg3812 Supplementary Text Figs. S1 to S15 Tables S1 and S2 References (44–54) 18. S. Takeda, T. Mizuta, M. Fuwa, P. van Loock, A. Furusawa, Nature 500, 315–318 (2013). Submitted 20 December 2022; accepted 19 April 2023 10.1126/science.adg3812 Sahu et al., Science 380, 718–721 (2023) 19 May 2023 4 of 4
10.1126_science.ade4884
RES EARCH CONSERVATION Widespread diversity deficits of coral reef sharks and rays Colin A. Simpfendorfer1,2*, Michael R. Heithaus3, Michelle R. Heupel2,4, M. Aaron MacNeil5, Mark Meekan6, Euan Harvey7, C. Samantha Sherman1,8, Leanne M. Currey-Randall4, Jordan S. Goetze9,10, Jeremy J. Kiszka3, Matthew J. Rees6,11, Conrad W. Speed6, Vinay Udyawer12, Mark E. Bond3, Kathryn I. Flowers3,13, Gina M. Clementi3, Jasmine Valentin-Albanese14, M. Shiham Adam15, Khadeeja Ali3,16, Jacob Asher17, Eva Aylagas17, Océane Beaufort18, Cecilie Benjamin19, Anthony T. F. Bernard20,21, Michael L. Berumen22, Stacy Bierwagen4, Chico Birrell23, Erika Bonnema3, Rosalind M. K. Bown24, Edward J. Brooks25, J. Jed Brown26, Dayne Buddo27, Patrick J. Burke28,29, Camila Cáceres3, Marta Cambra30,31, Diego Cardeñosa3, Jeffrey C. Carrier32, Sara Casareto3, Jennifer E. Caselle33, Venkatesh Charloo34, Joshua E. Cinner35, Thomas Claverie36, Eric E. G. Clua37,38, Jesse E. M. Cochran22, Neil Cook39,40, Jessica E. Cramp41,42, Brooke M. D’Alberto1,43, Martin de Graaf44, Mareike C. Dornhege45, Mario Espinoza30,31, Andy Estep46, Lanya Fanovich40, Naomi F. Farabaugh3, Daniel Fernando24, Carlos E. L. Ferreira47, Candace Y. A. Fields3,25, Anna L. Flam48, Camilla Floros49,50, Virginia Fourqurean51,52, Laura Gajdzik22,53, Laura García Barcia3, Ricardo Garla54,55, Kirk Gastrich3, Lachlan George2, Tommaso Giarrizzo56,57, Rory Graham5, Tristan L. Guttridge59,60, Valerie Hagan13, Royale S. Hardenstine16,22, Stephen M. Heck13, Aaron C. Henderson61, Patricia Heithaus3, Heidi Hertler61, Mauricio Hoyos Padilla62,63, Robert E. Hueter64,65, Rima W. Jabado1,66, Jean-Christophe Joyeux67, Vanessa Jaiteh68,69, Mohini Johnson70, Stacy D. Jupiter71, Muslimin Kaimuddin72,70, Devanshi Kasana3, Megan Kelley3, Steven T. Kessel73, Benedict Kiilu74, Taratau Kirata75†, Baraka Kuguru76, Fabian Kyne77, Tim Langlois78,79, Frida Lara80,81, Jaedon Lawe82, Elodie J. I. Lédée1, Steve Lindfield83, Andrea Luna-Acosta84, Jade Q. Maggs85, B. Mabel Manjaji-Matsumoto86, Andrea Marshall87, Lucy Martin88, Daniel Mateos-Molina89,90, Philip Matich60, Erin McCombs91, Ashlie McIvor22,92, Dianne McLean6,93, Llewelyn Meggs82, Stephen Moore1, Sushmita Mukherji1,2, Ryan Murray94, Stephen J. Newman95, Josep Nogués88, Clay Obota96,97, Domingo Ochavillo98, Owen O'Shea99,100, Kennedy E. Osuka96,101, Yannis P. Papastamatiou3, Nishan Perera23, Bradley Peterson14, Caio R. Pimentel67,102, Fabián Pina-Amargós103,104, Hudson T. Pinheiro105, Alessandro Ponzo106, Andhika Prasetyo107, L. M. Sjamsul Quamar108, Jessica R. Quinlan3, José Amorim Reis-Filho109, Hector Ruiz110, Alexei Ruiz-Abierno104, Enric Sala111, Pelayo Salinas-de-León112,113, Melita A. Samoilys96,114, William R. Sample3, Michelle Schärer-Umpierre110, Audrey M. Schlaff1, Kurt Schmid55,115, Sara N. Schoen3, Nikola Simpson116, Adam N. H. Smith117, Julia L. Y. Spaet118, Lauren Sparks119, Twan Stoffers120, Akshay Tanna24, Rubén Torres121, Michael J. Travers95, Maurits van Zinnicq Bergmann3,58, Laurent Vigliola122, Juney Ward123, Joseph D. Warren14, Alexandra M. Watts47,124, Colin K. Wen125, Elizabeth R. Whitman3, Aaron J. Wirsing126, Aljoscha Wothke40, Esteban Zarza-González127,128, Demian D. Chapman3,60 A global survey of coral reefs reveals that overfishing is driving resident shark species toward extinction, causing diversity deficits in reef elasmobranch (shark and ray) assemblages. Our species- level analysis revealed global declines of 60 to 73% for five common resident reef shark species and that individual shark species were not detected at 34 to 47% of surveyed reefs. As reefs become more shark-depleted, rays begin to dominate assemblages. Shark-dominated assemblages persist in wealthy nations with strong governance and in highly protected areas, whereas poverty, weak governance, and a lack of shark management are associated with depauperate assemblages mainly composed of rays. Without action to address these diversity deficits, loss of ecological function and ecosystem services will increasingly affect human communities. C oral reef ecosystems are under increas- ing pressure from human activities— including intense fishing, degraded water quality, and climate change (1, 2)—that threaten species supporting a wide range of ecosystem functions (3). Sharks and rays (hereafter “elasmobranchs”) have diverse roles Affiliations are listed at the end of this paper. *Corresponding author. Email: colin.simpfendorfer@jcu.edu.au †Deceased. on coral reefs as predators and prey across multiple trophic levels and in the cycling and movement of nutrients (3–5). Recent evidence indicates that overfishing has driven sharks toward functional extinction on many reefs. In a global survey, sharks were not observed on nearly 20% of reefs surveyed (6). Yet until re- cently, reef shark species were listed in lower risk extinction categories by the International Union for the Conservation of Nature (IUCN). With ~37% of all elasmobranch species threat- ened with extinction (7), a key question for coral reef ecosystems lies in understanding the global extent of species loss in elasmo- branch assemblages. We characterized elas- mobranch assemblage structure on coral reefs across a gradient of human pressures to esti- mate the local depletion and global extinction risk of the most common reef species, reveal- ing the human and environmental factors that influence assemblage structure and that lead to a deficit in predator diversity that could affect reef ecological functioning. To understand the extent of the reef elasmo- branch diversity deficit, we surveyed 391 coral reefs in 67 nations and territories using 22,756 baited remote underwater video stations (BRUVS). We examined reef-level species richness, spe- cies composition of elasmobranch assem- blages, and species relative abundance (MaxN; the maximum number of each species ob- served in a single frame of each 60-min de- ployment then averaged across all deployments on one reef) (8). We examined how elasmo- branch species assemblages changed in re- sponse to human pressures, using unweighted pair group with arithmetic mean (UPGMA) clustering to identify reefs with the most simi- lar assemblages (8). We then compared these clusters with estimated depletion of key resi- dent elasmobranch species at the reef level and examined whether socioeconomic, management, or environmental factors could predict cluster membership, using linear discriminant analy- sis. Reef-level depletion was estimated by divid- ing the observed mean MaxN of a species at individual reefs by a model-estimated baseline abundance (without human pressures) for each sampling site (a small group of closely asso- ciated reefs) and subtracting this value from 1. Baseline abundance (also expressed as MaxN) was estimated from a general linear model relating observed MaxN to sampling site, hu- man pressure [represented by total market gravity, the size and travel time to human mar- kets (2)], and marine protected area (MPA) status [closed to all fishing, open to fishing, or restricted (some fishing but with restric- tions)]. The baseline was estimated by setting all parameters to those expected at a site with no human pressure (gravity to the minimum for an ocean basin and protection status to closed) (8). Sampling identified 104 distinct elasmobranch species or species complexes (table S1), repre- senting more than 77% of elasmobranch spe- cies known to occur on coral reefs at some point during their lives (9). More than half (n = 53) of the species were rarely observed, with 10 or fewer sightings. We estimated reef- level depletion for the nine most commonly occurring species of shark [n = 5; Caribbean reef sharks (Carcharhinus perezi) and nurse sharks (Ginglymostoma cirratum) in the Atlantic; grey reef sharks (Carcharhinus amblyrhynchos), Simpfendorfer et al., Science 380, 1155–1160 (2023) 16 June 2023 1 of 6 A 30N 15N e d u t i t a L 0 15S 30S B 4 RES EARCH | R E S E A R C H A R T I C L E blacktip reef sharks (Carcharhinus melanopterus), and whitetip reef sharks (Triaenodon obesus) in the Indo-Pacific] and rays [n = 4; yellow stingrays (Urobatis jamaicensis) and southern stingrays (Hypanus americanus) in the Atlan- tic; blue spotted mask rays (Neotrygon spp.) and blue spotted ribbontail rays (Taeniura lymma and Taeniura lessoni) in the Indo- Pacific]. The Galapagos shark was excluded from estimates of global depletion because sampling only covered a relatively small pro- portion of its range, but the results for this species were broadly similar. The nine key resident species represented 77.7% of all elas- mobranchs observed in the study and are those that serve important ecological roles (10) and contribute the most to, and under- pin, livelihoods through fishing (11) and dive tourism (12). We found that mean depletion of five key resident reef sharks on individual reefs ranged from 100% depletion (none observed) to 0% (no depletion), averaging 62.8% (Fig. 1A). Mean depletion of key resident reef sharks followed the overall decline in elasmobranch abun- dance as measured with MaxN (Fig. 1B), de- creased as the fraction of the elasmobranch assemblage comprised of sharks decreased (Fig. 1C), and showed little change across a range of elasmobranch species richness (Fig. 1D); these patterns were generally consistent between ocean basins. Across the range of depletion, five main clusters of reefs were iden- tified in the Atlantic, and eight were identified in the Indo-Pacific (Figs. 2 and 3), including at least one cluster in each ocean basin (cluster 1 in the Atlantic and cluster 2 in the Indo- Pacific) having shark populations in a rela- tively intact state, with low levels of depletion of the five main resident reef shark species (Caribbean reef and nurse sharks in the At- lantic; grey reef, blacktip reef, and whitetip reef sharks in the Indo-Pacific) (8). Remain- ing clusters represented assemblages with increasing depletion of resident shark spe- cies and greater proportions of the overall elas- mobranch assemblage represented by rays (Figs. 2C and 3B). Both ocean basins show a similar transition through these assemblages as key resident shark species became depleted. The four key ray species (yellow and south- ern stingrays in the Atlantic; blue spotted mask and blue spotted ribbontail rays in the Indo-Pacific) increased only with depletion of one or more resident reef shark species, with rays dominating in the most shark-depleted areas. These predictable changes in assem- blage provide the ability to infer the status of reef shark populations, and the level of hu- man pressure they are experiencing, in future surveys. Elasmobranch species assemblage clusters on reefs in both basins were significantly relat- ed to certain socioeconomic and manage- Depletion 1.00 0.75 0.50 0.25 0.00 30E 60E 90E 120E 150E 180E Longitude 150W 120W 90W 60W 30W Atlantic Indo−Pacific 3 ) Basin p o r d r e p N x a M h c n a r b o m s a e ( 2 l e c n a d n u b A 1 0 C k r a h s d e s i r p m o c N x a M h c n a r b o m s a e l . p o r P 1.00 0.75 0.50 0.25 0.00 D 25 20 s s e n h c i r i s e c e p S 15 10 5 0 1.00 0.75 0.25 0.50 Shark depletion 0.00 1.00 0.50 0.75 0.25 Shark depletion 0.00 1.00 0.75 0.50 0.25 Shark depletion 0.00 Fig. 1. The global decline of coral reef elasmobranchs. (A) Reef-scale estimates of depletion of resident coral reef shark species. Depletion is proportion of unfished population lost, represented as the measured MaxN as a proportion of MaxN in an unfished state (gravity, lowest in basin; MPA status, closed) (8). Open circles indicate no sharks or rays were observed; gray circles indicate none of the resident shark species used to calculate mean depletion were present. (B) Relationship between depletion of resident shark species and MaxN by ocean basin. (C) Relationship between depletion of resident shark species and the proportion of elasmobranch MaxN that comprised shark, demonstrating the transition from shark- to ray-dominated assemblages. (D) Relationship between depletion of resident shark species and species richness. ment factors, with linear discriminant analysis (LDA) accounting for ~85% of variance be- tween clusters (tables S2 and S3). Important socioeconomic factors included the Human Development Index (an index of a nation’s level of education, life expectancy and stan- dard of living) and Voice and Accountability Index (an index of the extent to which people in each nation can participate in governance, free expression, free media, and free associ- ation). Important management factors were whether the reef occurred in a marine pro- tected area (MPA) or whether a reef was within a nation where all targeted shark fishing and trade is prohibited, known as a “shark sanc- tuary.” Given that shark sanctuaries have largely been implemented in nations in which fishing for sharks was limited for economic or cultural reasons (6), their effectiveness as tools for recovering reef shark populations remains an open question. Total market grav- ity was more important in the Indo-Pacific than the Atlantic, possibly because remote reefs (>4 hours travel time from human settlements) Simpfendorfer et al., Science 380, 1155–1160 (2023) 16 June 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Structure of shark and ray assemblages on Atlantic coral reefs. (A and B) Clusters of reefs with similar species composition from UPGMA clustering of 106 reefs in the Atlantic basin based on a global set of 31 coral reef– associated species. Five main clusters, representing 87.0% of reefs, were identified. Their locations are indicated with colored triangles. Reefs with minor clusters are indicated with gray dots (n = 7). Reefs where no elasmobranchs were observed are indicated with black dots (n = 5). (C) Regime plot showing all species assemblage clusters as a function of the mean depletion of the resident reef shark species (Caribbean reef and nurse sharks) and the proportion of all observed elasmobranchs that were sharks. Size of points (and numbers) indicate the number of reefs in each cluster, and colors indicate cluster identity as per (A). (D) Population level relative to original levels of four resident reef species in each of the five main clusters. Proportion of original level = 1 – depletion. Horizontal lines indicate mean, boxes indicates 25 to 75 percentile, and whiskers indicate 95% confidence interval. A 30N e d u t i t a L 20N 10N C k r a h s f o n o i t r o p o r P 1.00 0.75 0.50 0.25 0.00 90W 80W 70W 60W B 45W 30W Cluster 1 2 6 10 3 Brazil 0 30N 10S 20N 20S 10N 0 10S 20S 90W 80W Longitude 70W 1.00 0.75 0.50 0.25 0.00 60W D 45W Longitude 30W Caribbean reef shark 1 12 25 2 14 6 6 2 1.00 0.75 0.50 Shark depletion 0.25 1 1.00 0.75 0.50 0.25 0.00 0.00 1.0 0.5 0.0 1.0 0.5 0.0 4 2 0 4 2 0 l i a n g i r o n o i t r o p o r P 1 3 Nurse shark 2 6 10 1 3 2 6 10 Southern stingray 1 3 2 6 10 Yellow stingray 1 3 2 6 10 are relatively rare in the Atlantic compared with the Indo-Pacific (fig. S1) (13). Environ- mental factors (coral cover and relief) had lit- tle influence in predicting cluster membership. Elasmobranch assemblage structure on coral reefs in both the Atlantic and Indo-Pacific are therefore mainly driven by management and socioeconomic factors, with shark-dominated assemblages more likely to occur in wealthy, well-governed nations and in highly protected areas or shark sanctuaries, whereas poverty, limited governance, and a lack of shark pro- tection are associated with assemblages mainly composed of rays. To further characterize the diversity deficits that underpin these assemblage differences, we compared species observations in our BRUVS with their historical ranges drawn from pub- lished literature, including historical accounts, and found that sharks were not detected at 13.6% of reefs (19 Atlantic and 34 Indo-Pacific), whereas rays were not detected at 21.5% of reefs (10 Atlantic and 74 Indo-Pacific); both groups were not detected at 6.6% of reefs sur- veyed (5 Atlantic and 19 Indo-Pacific). At the species level, absences were severe. On the basis of their known historic distribution, def- icits were 46.9% of reefs (112 of 246) for black- tip reef sharks, 41.3% (31 of 75) for Caribbean reef sharks, 40.8% (102 of 250) for grey reef sharks, 36.2% (89 of 246) for whitetip reef sharks, and 34.7% (n = 26 of 75) for nurse sharks (fig. S2). Among rays, deficits were even more stark: 78.9% (75 of 95) for yellow stingray, 62.8% (81 of 129) for blue spotted ribbontail rays, and 55.6% (79 of 142) for blue spotted maskrays. An exception was the southern sting- ray, which was not detected at only 19.8% (n = 20 of 101) of expected reefs in the Atlantic. A failure to detect rays may not always indicate absence because they are often cryptic and therefore missed on BRUVS, especially when sharks are present (14). Collectively, these di- versity deficits show that elasmobranch loss on coral reefs is more extensive than previ- ously demonstrated, with widespread losses of key species across many of the world’s coral reefs, especially in Asia, eastern Africa, conti- nental South America, and the central-eastern Caribbean. Previous estimates of the status of reef shark and ray species have been geographically limited, varying among surveyed reefs from very high abundances (15) to local extinction (16). This disparity has made it difficult to assess the global status of individual species. Therefore, we used our estimates of reef-level depletion to estimate the global depletion and extinction risk of the most common res- ident reef sharks (five species) and rays (four species). Mean and standard error reef-level depletion was calculated within jurisdic- tions (nations or remote territories) and used to produce confidence intervals for jurisdic- tional depletion levels. To estimate an overall global depletion level by species, we weighted the jurisdictional depletion by the percent- age of the world’s coral reefs in their waters and produced a weighted global mean de- pletion (8). Extinction risk was estimated by comparing proportional global depletion to the criteria for the IUCN Red List A2 (pop- ulation decline) category (17), assuming that the decline had occurred in the past three generations (29 to 90 years). In IUCN assess- ments before the availability of this global survey, all reef-resident shark species were considered at lower risk of extinction (Near Threatened) (18). Grey reef shark had the highest level of global decline [69.8% ± 1 stan- dard error (SE) 62.6 to 77.1], followed by nurse shark (68.6% ± 49.7 to 87.4), Caribbean reef shark (64.8% ± 42.0 to 87.5), blacktip reef shark (64.5% ± 58.7 to 70.4), and white- tip reef shark (60.4% ± 51.2 to 70.2) (Fig. 4). The estimated declines of resident species of reef sharks met the IUCN Red List criteria for Endangered. Population changes of rays were more variable, with increasing populations in some nations and declines in others (fig. S3), reflecting the compositional changes seen across our gradient of human pressures. When examined at the global level, no ray species ex- amined met criteria for elevated extinction risk, which is consistent with current nonthreatened status of these species on the Red List. Simpfendorfer et al., Science 380, 1155–1160 (2023) 16 June 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E 30E 60E 90E 120E 150E 180E 150W 120W 90W A 30N 15N e d u t i t a L 0 15S 30S B 1.00 30E 60E 90E 120E 1.00 0.75 0.50 0.25 0.00 1.00 7 2 85 1 57 6 27 0.75 1 3 k r a h s 0.50 f o n o i t r o p o r P 0.25 0.00 2 4 12 3 22 1 1 6 1.00 0.50 0.75 Shark depletion 0.25 0.75 0.50 0.25 0.00 0.00 30N 15N 0 15S 30S Cluster 1 2 5 7 10 13 14 18 150W 120W 90W 180E 150E Longitude C Grey reef shark 1.0 0.5 0.0 1.0 0.5 0.0 1.0 0.5 0.0 2 0 3 2 1 0 l i a n g i r o n o i t r o p o r P 2 5 1 10 13 14 18 7 Blacktip reef shark 2 5 1 10 13 14 18 7 Whitetip reef shark 2 5 1 10 13 14 18 7 Blue spotted maskray 2 5 1 10 13 14 18 7 Blue spotted ribbontail ray 2 5 1 10 13 14 18 7 Fig. 3. Structure of shark and ray assemblages on Indo-Pacific coral reefs. (A) Clusters of reefs with similar species composition from UPGMA clustering of 285 reefs in the Indo-Pacific basin based on a global set of 31 coral reef- associated species. Eight main clusters, representing 82.1% of reefs, were identified. Their locations are indicated with colored triangles. Reefs with minor clusters are indicated with gray dots (n = 30). Reefs where no elasmobranchs were observed are indicated with black dots (n = 21). (B) Regime plot showing all species assemblage clusters as a function of the mean depletion of the resident species of reef shark (grey reef, blacktip reef, whitetip reef, and Galapagos sharks) and the proportion of all observed elasmobranchs that were sharks. Size of points (and numbers) indicate the number of reefs in each cluster, and colors indicate cluster identity as per (A); minor clusters are indicated in gray. (C) Population level relative to original levels of five core shark and ray species in each of the eight main species assemblage clusters. Proportion of original level = 1 – depletion. Horizontal lines indicate mean, boxes indicate 25 to 75 percentile, and whiskers indicate 95% confidence interval. Our study of nations hosting ~90% of global reefs reveals that resident reef shark species are at much higher risk of extinction than previously thought. Local declines, shaped by human pressures that vary across ocean basins, have led to consistent changes in the structure of coral reef elasmobranch assem- blages that may have profound effects on the broader ecosystem. The direct and indi- rect effects of fishing have driven shifts in species composition from shark-dominated to ray-dominated assemblages and ultimate- ly the complete loss of sharks and rays at a small proportion (~ 7%) of reefs surveyed. In addition to changes in the structure of as- semblages, all major resident shark species have declined to such levels that they qualify as Endangered by the IUCN Red List Criteria. These changes wrought on coral reef elasmo- branch assemblages demonstrate the per- vasiveness of fishing on coral reefs (19) and the substantial risks to reef-dependent hu- man communities of continued overfishing. Elasmobranch species vary widely in their eco- nomic value, with some fished for subsistence, others fished for local or export markets, and others valued alive as tourism resources (12, 20). Thus, understanding threats and conservation options for rebuilding populations at a species level will assist in developing effective man- agement of coral reef elasmobranchs as part of a sustainable social-ecological system. Although reef sharks are at considerable risk over broad spatial scales, our results show that declines at one reef will have little effect on reefs tens to hundreds of kilometers dis- tant. Thus, despite populations being func- tionally extinct at the reef level, the potential to rebuild abundances remains relatively high if there are protected areas or strong fisheries management within a region (6). These source populations are present among many small oceanic islands where low human populations and the high cultural value of sharks has re- sulted in fishing levels that are below those seen elsewhere (21). MPAs also provide the opportunity to act as source populations; how- ever, their designation alone is insufficient to deliver benefits. As others have observed (22), high compliance is required. We show that there are reefs in regions with widespread depletion of reef shark species that had metrics indicating that they are in a relatively healthy state compared with those around them. These Simpfendorfer et al., Science 380, 1155–1160 (2023) 16 June 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E A n o i t a N Sri Lanka Kenya Qatar Japan Indonesia Tanzania Taiwan Saudi Arabia (Red Sea) Philippines Vanuatu USA−Pacific Madagascar South Africa Mayotte Mozambique Christmas Island American Samoa Malaysia Guam Samoa Species Blacktip reef Grey reef Whitetip reef Papua New Guinea Seychelles Cook Islands Maldives French Polynesia Australia−Indian Ocean Fiji Kiribati Fed. States of Micronesia Niue Cocos−Keeling Palau Solomon Islands Australia−Pacific New Caledonia Tonga Jarvis Island Species Caribbean reef Nurse B Martinique Jamaica Montserrat Dominican Republic Barbados Colombia USA−Western Atlantic Trinidad and Tobago Brazil Cuba Belize Pedro Bank The Bahamas Turks and Caicos Colombia (Offshore Islands) Dutch Caribbean South Antigua and Barbuda 1.00 0.75 0.50 Depletion level 0.25 0.00 1.00 0.75 0.50 Depletion level 0.25 0.00 Fig. 4. Depletion of core coral reef shark species in the Indo-Pacific and Atlantic basins at national or near-national scale. (A) Indo-Pacific basin. (B) Atlantic basin. Depletion was calculated by comparing reef-level species MaxN values to unfished, estimated by using a linear model in which market gravity (a measure of the human pressure from population and access to reefs) was set to the ocean basin minimum and reef protected status was “closed” (no take MPA) (8). Reef-level depletion scores were modeled by nation and used to estimate a global level of depletion (vertical dashed lines) ± 1 standard error (shaded area) calculated by weighting national-level depletion by coral reef area (as a percent of global total coral reef area that occurs within the range of each shark species). included Tubbataha (Philippines), Sipidan (Island Malaysia), Glover’s Reef and Light- house Reef (Belize), and Misool (Indonesia); in all of these locations, there are programs to actively manage and enforce MPA regulations that are likely to account for these successes (23–25). Multiple nations have strong management measures (such as spatial protections and/or fishing restrictions) in place that benefit reef species. This study builds the case that species-specific reef shark management pro- vides the best way forward for conservation and rebuilding of reef sharks in places where they have declined, among nations with the desire and capacity to do so (7, 8). Recent studies show that populations of reef sharks can rebound in under a decade if appropri- ate management strategies that reduce fish- ing pressure are in place (26). Although direct management is critical, local and national socioeconomic factors that affect the ability of nations to develop, implement, and enforce regulations, and the likelihood that fishers comply with regulations, will be critical to maintaining or rebuilding populations and diverse elasmobranch assemblages. If not ad- dressed, pressures causing the shark and ray diversity deficits we outline will continue to result in a loss of species, ecological func- tions, and ecosystem services that support sustainable livelihoods for millions of people worldwide. Simpfendorfer et al., Science 380, 1155–1160 (2023) 16 June 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E RE FE RENCES AND N OT ES 1. T. P. Hughes, D. R. Bellwood, S. R. Connolly, H. V. Cornell, 2. R. H. Karlson, Curr. Biol. 24, 2946–2951 (2014). J. E. Cinner et al., Proc. Natl. Acad. Sci. U.S.A. 115, E6116–E6125 (2018). 3. G. Roff et al., Trends Ecol. Evol. 31, 395–407 (2016). 4. K. I. Flowers, M. R. Heithaus, Y. P. Papastamatiou, Fish Fish. 5. 22, 105–127 (2021). J. J. Williams, Y. P. Papastamatiou, J. E. Caselle, D. Bradley, D. M. P. Jacoby, Mobile marine predators: an understudied source of nutrients to coral reefs in an unfished atoll. Proc. R. Soc. B Biol. Sci. 10.1098/rspb.2017.2456 (2018). 6. M. A. MacNeil et al., Nature 583, 801–806 (2020). 7. N. K. Dulvy et al., Curr. Biol. 31, 4773–4787.e8 (2021). 8. Materials and methods are available as supplementary materials. 9. C. S. Sherman et al., Nat. Commun. 14, 15 (2023). 10. M. R. Heupel, Y. P. Papastamatiou, M. Espinoza, M. E. Green, C. A. Simpfendorfer, Front. Mar. Sci. 6, 12 (2019). 11. S. A. Appleyard, W. T. White, S. Vieira, B. Sabub, Sci. Rep. 8, 6693 (2018). 12. A. M. Cisneros-Montemayor, M. Barnes-Mauthe, D. Al-Abdulrazzak, E. Navarro-Holm, U. R. Sumaila, Oryx 47, 381–388 (2013). 13. T. R. McClanahan, Mar. Policy 119, 104022 (2020). 14. C. S. Sherman, M. R. Heupel, S. K. Moore, A. Chin, C. A. Simpfendorfer, Mar. Ecol. Prog. Ser. 641, 145–157 (2020). 15. J. Mourier et al., Curr. Biol. 26, 2011–2016 (2016). 16. M. O. Nadon et al., Conserv. Biol. 26, 493–503 (2012). IUCN, IUCN Red List categories and criteria, version 3.1 17. (IUCN, ed. 2, 2012). 18. N. K. Dulvy et al., Aquat. Conserv. 26, 134–153 (2016). 19. T. D. Eddy et al., One Earth 4, 1278–1285 (2021). 20. H. Booth, D. Squires, E. J. Milner-Gulland, Ocean Coast. Manage. 182, 104994 (2019). 21. J. S. Goetze et al., PLOS ONE 13, e0200960 (2018). 22. G. J. Edgar et al., Nature 506, 216–220 (2014). 23. R. Murray et al., J. Asia-Pac. Biodivers. 12, 49–56 (2019). 24. V. F. Jaiteh et al., Front. Mar. Sci. 3, (2016). 25. G. Clementi et al., Mar. Ecol. Prog. Ser. 661, 175–186 (2021). 26. C. W. Speed, M. Cappo, M. G. Meekan, Biol. Conserv. 220, 308–319 (2018). 27. C. Simpfendorfer, Widespread diversity deficits of coral reef sharks and rays. Dryad (2023); https://doi.org/10.5061/ dryad.qbzkh18h0. 28. C. Simpfendorfer, Widespread diversity deficits of coral reef sharks and rays. Zenodo (2023); https://doi.org/10.5281/ zenodo.7030578. ACKN OW LEDG MEN TS We thank our individual funders for country-specific deployments, whose contributions greatly enhanced the sampling coverage of the projects; all of the government permitting agencies that allowed us to work in their waters; and the Global FinPrint volunteers from Stony Brook University, Florida International University, James Cook University, the Aquarium of the Pacific, and Shedd Aquarium who watched the BRUVS footage. Funding: Core funding for Global FinPrint was provided by the Paul G. Allen Family Foundation (to D.D.C. and M.R.Hei.). Author contributions: Conceptualization: D.D.C., M.R.Hei., C.A.S., M.R.Heu., M.A.M., M.M., and E.H. Methodology: D.D.C., M.R.Hei., C.A.S., M.R.Heu., M.A.M., M.M., and E.H. Investigation: All authors. Visualization: C.A.S. Funding acquisition: D.D.C. and M.R.Hei. Project administration: D.D.C., M.R.Hei., C.A.S., M.R.Heu., M.A.M., M.M., and E.H. Writing – original draft: C.A.S., D.D.C., M.R.Hei., M.R.Heu., M.A.M., M.M., E.H., and C.S.S. Writing – review and editing: All authors. Competing interests: The authors declare that they have no competing interests Data and materials availability: Data files have been deposited in Dryad (27), and R script has been deposited in Zenodo (28). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse 1College of Science and Engineering, James Cook University, Townsville, QLD, Australia. 2Institute for Marine and Antarctic Studies, University of Tasmania, Hobart, TAS, Australia. 3Institute of Environment, Department of Biological Sciences, Florida International University, North Miami, FL, USA. 4Australian Institute of Marine Science, Townsville, QLD, Australia. 5Ocean Frontier Institute, Department of Biology, Dalhousie University, Halifax, NS, Canada. 6Australian Institute of Marine Science, Perth, WA, Australia. 7School of Molecular and Life Sciences, Curtin University, Bentley, WA, Australia. 8Earth to Ocean Group, Biological Sciences, Simon Fraser University, Burnaby, BC, Canada. 9School of Molecular and Life Sciences, Curtin University, Perth, WA, Australia. 10Marine Science Program, Biodiversity and Conservation Science, Department of Biodiversity, Conservation and Attractions, Perth, WA, Australia. 11Centre for Sustainable Ecosystems Solutions, School of Earth, Atmospheric and Life Sciences, University of Wollongong, Wollongong, NSW, Australia. 12Australian Institute of Marine Science, Darwin, NT, Australia. 13Sharks and Rays Conservation Program, Mote Marine Laboratory, Sarasota, FL, USA. 14School of Marine and Atmospheric Sciences, Stony Brook University, Stony Brook, NY, USA. 15International Pole and Line Foundation–Maldives, Malé, Republic of Maldives. 16Maldives Marine Research Institute, Ministry of Fisheries, Marine Resources and Agriculture, Malé, Republic of Maldives. 17Red Sea Global, Department of Environmental Protection and Regeneration, AlRaidah Digital City, Riyadh, Saudi Arabia. 18Kap Natirel NGO, Fort l’Olive, Guadeloupe, France. 19Mahonia Na Dari Research and Conservation Centre, Kimbe, Papua New Guinea. 20South African Institute for Aquatic Biodiversity, National Research Foundation, Makhanda, South Africa. 21Department of Zoology and Entomology, Rhodes University, Makhanda, South Africa. 22Red Sea Research Center, King Abdullah University of Science and Technology, Thuwal, Saudi Arabia. 23Marine Conservation, Madagascar Program, Wildlife Conservation Society, Antananarivo, Madagascar. 24Blue Resources Trust, Colombo, Sri Lanka. 25Cape Eleuthera Institute, Cape Eleuthera, Eleuthera, The Bahamas. 26Center for Sustainable Development, College of Arts and Sciences, Qatar University, Doha, Qatar. 27Georgia Aquarium–IUCN Center for Species Survival, Atlanta, GA, USA. 28School of Natural Sciences, Macquarie University, Sydney, NSW, Australia. 29Bimini Biological Field Station, Bimini, Bahama. 30Centro de Investigación en Ciencias del Mar y Limnología, Universidad de Costa Rica, San José, Costa Rica. 31MigraMar, Olema, CA, USA. 32Department of Biology, Albion College, Albion, MI, USA. 33Marine Science Institute, University of California Santa Barbara, Santa Barbara, CA, USA. 34Coastal Impact, Goa, India. 35College of Arts, Society, and Education, James Cook University, Townsville, QLD, Australia. 36Centre Universitaire de Formation et de Recherche de Mayotte, Dembeni, France. 37Paris Sciences Lettres, Centre de Recherche Insulaire et Observatoire de l’Environnement, Opunohu Bay, Papetoai, French Polynesia. 38Laboratoires d’Excellence Corail, Ecole Pratique des Hautes Etudes, Perpignan, France. 39School of Biosciences, Cardiff University, Cardiff, UK. 40Environmental Research Institute Charlotte- ville, Charlotteville, Trinidad and Tobago. 41Australian Research Council Centre of Excellence for Coral Reef Studies, James Cook University, Townsville, QLD, Australia. 42Sharks Pacific, Rarotonga, Cook Islands. 43Oceans and Atmosphere, Commonwealth Scientific and Industrial Research Organization, Hobart, TAS, Australia. 44Wageningen Marine Research, Wageningen University & Research, IJmuiden, Netherlands. 45Graduate School for Global Environmental Studies, Sophia University, Tokyo, Japan. 46Waitt Institute, La Jolla, CA, USA. 47Reef Systems Ecology and Conservation Lab, Departamento de Biologia Marinha, Universidade Federal Fluminense, Rio de Janeiro, Brazil. 48Marine Megafauna Foundation, Palm Beach, FL, USA. 49Oceanographic Research Institute, Durban, South Africa. 50TRAFFIC International, Cambridge, UK. 51College of Arts, Science, and Education, Florida International University, North Miami, FL, USA. 52Science Department, Georgia Jones-Ayers Middle School, Miami, FL, USA. 53Division of Aquatic Resources, Department of Land and Natural Resources, Honolulu, HI, USA. 54Centro de Biociências, Departmento de Botânica e Zoologia, Universidade Federal do Rio Grande do Norte, Brazil. 55Beacon Development Company, King Abdullah University of Science and Technology, Thuwal, Saudi Arabia. 56Instituto de Ciencias do Mar, Universidade Federal do Ceará, Fortaleza, Brazil. 57Grupo de Ecologia Aquática, Espaço Inovação do Parque de Ciência e Tecnologia Guamá, Guamá, Pará, Brazil. 58Independent consultant, Hull, UK. 59Bimini Biological Field Station Foundation, South Bimini, The Bahamas. 60Saving the Blue, Cooper City, FL, USA. 61The School for Field Studies, Center for Marine Resource Studies, South Caicos, Turks and Caicos Islands. 62Pelagios Kakunjá, La Paz, Mexico. 63Fins Attached, Colorado Springs, CO, USA. 64Center for Shark Research, Mote Marine Laboratory, Sarasota, FL, USA. 65OCEARCH, Park City, UT, USA. 66Elasmo Project, Dubai, United Arab Emirates. 67Departamento de Oceanografia e Ecologia, Universidade Federal do Espírito Santo, Vitória, Espírito Santo, Brazil. 68Murdoch University, Murdoch, WA, Australia. 69Centre for Development and Environment, University of Bern, Bern, Switzerland. 70Operation Wallacea, Spilsby, Lincolnshire, UK. 71Melanesia Program, Wildlife Conservation Society, Suva, Fiji. 72Wasage Divers, Wakatobi & Buton, Southeast Sulawesi, Indonesia. 73Daniel P. Haerther Center for Conservation and Research, John G. Shedd Aquarium, Chicago, IL, USA. 74Kenya Fisheries Service, Mombasa, Kenya. 75Ministry of Fisheries and Marine Resources, Kiritimati, Kiribati. 76Tanzania Fisheries Research Institute, Dar Es Salaam, Tanzania. 77University of the West Indies, Kingston, Jamaica. 78School of Biological Sciences, University of Western Australia, Perth, WA, Australia. 79The UWA Oceans Institute, University of Western Australia, Perth, WA, Australia. 80Departamento de Pesquerias, Centro Interdisciplinario de Ciencias Marinas del IPN, La Paz, Baja California Sur, Mexico. 81Pelagios Kakunjá, La Paz, Baja California Sur, Mexico. 82Yardie Environmental Conservationists Limited, Kingston, Jamaica. 83Coral Reef Research Foundation, Koror, Palau. 84Departamento de Ecología y Territorio, Facultad de Estudios Ambientales y Rurales, Pontificia Universidad Javeriana, Bogotá, Colombia. 85National Institute of Water and Atmospheric Research, Auckland, New Zealand. 86Borneo Marine Research Institute, Universiti Malaysia Sabah, Kota Kinabalu, Sabah, Malaysia. 87Marine Megafauna Foundation, West Palm, FL, USA. 88Island Conservation Society Seychelles, Victoria, Mahé, Seychelles. 89Emirates Nature - World Wide Fund for Nature, Dubai, United Arab Emirates. 90College of Marine Sciences and Aquatic Biology, University of Khorfakkan, Sharjah, UAE. 91Aquarium of the Pacific, Long Beach, CA, USA. 92Marine and Environmental Sciences Centre/ Aquatic Research Network, Regional Agency for the Development of Research, Technology and Innovation, Funchal, Madeira, Portugal. 93Oceans Institute, University of Western Australia, Perth, WA, Australia. 94Inland Fisheries Ireland, Dublin, Ireland. 95Western Australian Fisheries and Marine Research Laboratories, Department of Primary Industries and Regional Development, Government of Western Australia, Hillarys, WA, Australia. 96CORDIO East Africa, Mombasa, Kenya. 97Blue Ventures, Mombasa, Kenya. 98American Samoa Department of Marine and Wildlife Resources, Pago Pago, American Samoa. 99The Centre for Ocean Research and Education, Gregory Town, Eleuthera, The Bahamas. 100Department of Ocean Science, Memorial University, NL, Canada. 101Department of Environment and Geography, University of York, York, UK. 102Departamento de Ciências Agrárias e Biológicas, Universidade Federal do Espírito Santo, São Mateus, Espírito Santo, Brazil. 103Blue Sanctuary-Avalon, Jardines de la Reina, Cuba. 104Centro de Investigaciones Marinas, Universidad de La Habana, Habana, Cuba. 105Center for Marine Biology, University of São Paulo, São Sebastião, São Paulo, Brazil. 106Large Marine Vertebrates Research Institute Philippines, Puerto Princesa City, Palawan, Philippines. 107Center for Fisheries Research, Ministry for Marine Affairs and Fisheries, Jakarta Utara, Indonesia. 108Fisheries Department, Universitas Dayanu Ikhsanuddin, Bau Bau, Southeast Sulawesi, Indonesia. 109Programa de Pós Graduação em Ecologia: Teoria, Aplicação e Valores, Instituto de Biologia, Universidade Federal da Bahia, Salvador, BA, Brazil. 110HJR Reefscaping, Boquerón, Puerto Rico. 111Pristine Seas, National Geographic Society, Washington, DC, USA. 112Charles Darwin Research Station, Charles Darwin Foundation, Puerto Ayora, Galapagos Islands, Ecuador. 113Save Our Seas Foundation Shark Research Center and Guy Harvey Research Institute, Nova Southeastern University, Dania Beach, FL, USA. 114School of Pure and Applied Sciences, Pwani University, Kilifi, Kenya. 115Thurgau Hunting and Fishing Administration, Frauenfeld, Switzerland. 116SalvageBlue, Kingstown, Saint Vincent and the Grenadines. 117School of Mathematical and Computational Sciences, Massey University, Auckland, New Zealand. 118Evolutionary Ecology Group, Department of Zoology, University of Cambridge, Cambridge, UK. 119Indo Ocean Project, Jln Toyapakeh DESA Toyapakeh, Nusa Penida, Bali, Indonesia. 120Aquaculture and Fisheries Group, Wageningen University & Research, Wageningen, Netherlands. 121Reef Check Dominican Republic, Santo Domingo, Dominican Republic. 122Institut de Recherche pour le Développement, UMR Entropie (IRD- UR-UNC-CNRS-IFREMER), Nouméa, New Caledonia, France. 123Secretariat of the Pacific Regional Environment Programme, Apia, Samoa. 124Department of Natural Sciences, Faculty of Science Engineering, Manchester Metropolitan University, Manchester, UK. 125Department of Life Science, Tunghai University, Taichung, Taiwan. 126School of Environmental and Forest Sciences, University of Washington, Seattle, WA, USA. 127GIBEAM Research Group, Universi- dad del Sinú, Cartagena, Colombia. 128Corales del Rosario and San Bernardo National Natural Park, Colombia. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.ade4884 Materials and Methods Figs. S1 to S8 Tables S1 to S9 References (29–37) MDAR Reproducibility Checklist Submitted 31 August 2022; accepted 27 April 2023 10.1126/science.ade4884 Simpfendorfer et al., Science 380, 1155–1160 (2023) 16 June 2023 6 of 6
10.1126_science.adg4421
RES EARCH CHEMICAL PHYSICS Femtosecond symmetry breaking and coherent relaxation of methane cations via x-ray spectroscopy Enrico Ridente1,2†, Diptarka Hait1,2†‡, Eric A. Haugen1,2, Andrew D. Ross1,2§, Daniel M. Neumark1,2, Martin Head-Gordon1,2, Stephen R. Leone1,2,3* Understanding the relaxation pathways of photoexcited molecules is essential to gain atomistic-level insight into photochemistry. We performed a time-resolved study of ultrafast molecular symmetry breaking through geometric relaxation (Jahn-Teller distortion) on the methane cation. Attosecond transient absorption spectroscopy with soft x-rays at the carbon K-edge revealed that the distortion occurred within 10 ± 2 femtoseconds after few-femtosecond strong-field ionization of methane. The distortion activated coherent oscillations in the asymmetric scissoring vibrational mode of the symmetry-broken cation, which were detected in the x-ray signal. These oscillations were damped within 58 ± 13 femtoseconds because vibrational coherence was lost with the energy redistributing into lower-frequency vibrational modes. This study completely reconstructs the molecular relaxation dynamics of this prototypical example and opens avenues for exploring complex systems. C hemical reactions arise from the motion of atomic nuclei. Atomic displacements can be described in terms of either fluc- tuations about a local minimum of energy or relaxation toward such a minimum from a nonequilibrium configuration. The latter often results from interaction with light because photon absorption can lead to excited electronic states with minimum energy geometries quite distinct from the initial starting point. The nonequilibrium configurations arising from light-matter interaction thus can have sub- stantial surplus potential energy, which can drive chemical transformations. Therefore, intramolecular relaxation dynamics of photo- excited molecules are of fundamental photo- chemical interest. Jahn-Teller (JT) distortion (1, 2) is a special type of relaxation mechanism that spontane- ously reduces the spatial symmetry of non- linear molecules in degenerate electronic states. Molecular geometries where multiple electronic states are isoenergetic are not sta- ble for any of the associated states (1) and represent a fundamental breakdown of the Born-Oppenheimer approximation (3). It there- fore becomes energetically favorable to under- go distortions that lift the degeneracy by breaking spatial symmetry. JT distortions are ubiquitous in solids (4) and gas-phase mol- ecules (5). In this work, we used attosecond x-ray transient absorption spectroscopy (XTAS) to study symmetry breaking of the methane +) generated through vertical cation (CH4 1Department of Chemistry, University of California, Berkeley, CA 94720, USA. 2Chemical Sciences Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA. 3Department of Physics, University of California, Berkeley, CA 94720, USA. *Corresponding author. Email: srl@berkeley.edu †These authors contributed equally to this work. ‡Present address: Department of Chemistry and PULSE Institute, Stanford University, Stanford, CA 94305, USA. §Present address: TOPTICA Photonics Inc., Pittsford, NY 14534, USA. strong-field ionization (SFI). This work un- equivocally revealed the role and timescale of JT-induced dynamics experimentally and pro- vided an understanding of relaxation mecha- nisms in molecular systems. 21t2 21t2 CH4 22a1 22a1 + is a classic system in which JT dis- tortions occur (6–9). The process starts with CH4, which is the smallest stable molecule with tetrahedral (Td) geometry. The equilibrium C-H bond distances are 1.087 Å (10), and all of the H-C-H bond angles are ≈109.5° because of Td symmetry. The ground-state molecular orbitals (MOs) are shown in Fig. 1A, and the electronic configuration of neutral CH4 is 6 (1A1). The electronic ground-state 1a1 + at the Td geometry is configuration of CH4 5, which is triply degenerate (2T2) 1a1 because each of the three 1t2 orbitals is equally + likely to be singly occupied. Therefore, CH4 undergoes JT distortion away from the Td geometry to a lower symmetry C2v form (7, 11–14). This JT distortion involves t2 and e symmetry + the vibrational modes of CH4, making CH4 simplest T2 (cid:1) (t2 + e) JT problem (15). The resulting C2v equilibrium structure was com- puted to have two long (1.187 Å) and two short (1.083 Å) C-H bonds (Fig. 1A), which indicates antisymmetric stretching (t2 symmetry) relative to neutral CH4. The angle formed by the long C-H bonds is 55.0°, and the short bonds form an angle of 125.7°, representing considerable deviations from the initial Td geometry through bending motions (of t2 and e symmetry in CH4). These distortions lower the energy of the doubly occupied 3a1 and 1b1 MOs (Fig. 1A) but also destabilize the 1b2 singly occupied MO (SOMO). The electronic ground + therefore is 2B2 (12). state of CH4 + in Molecular JT-distorted forms, and CH4 particular, have been extensively studied both theoretically (15–18) and experimentally (6, 19). Experimentally, time resolving the JT distor- + remained an open challenge tion in CH4 (7, 13) because of the ultrafast nature of the process. JT-distorted species have been charac- terized in photoelectron spectroscopy experi- ments (6, 20), but such measurements lack the temporal resolution to obtain the femto- second timescale dynamics of symmetry break- ing. Baker et al. (21) used attosecond-resolution high-harmonic emission spectroscopy to report + and on the onset of the JT distortions in CH4 + up to the first 1.6 fs. The nu- deuterated CD4 clear motion in those experiments, however, cannot be reconstructed at longer times, which precludes a complete analysis of the JT rela- xation process and subsequent coherent mo- tion. Coulomb explosion experiments by Li et al. (22) probed the dynamics of CH4 + by recording the photofragments after inter- action with two time-delayed strong-field 800-nm pulses, but temporal resolution was limited by the 25-fs pulses used as pump and probe. Furthermore, their use of a multicycle pump pulse led to several additional photo- products from higher-energy fragmentation pathways that compete with JT distortions. It has been shown that the use of shorter, few- cycle 800-nm pulses increased the relative + by suppressing additional pro- amount of CH4 duct channels (23). XTAS, based on attosecond- and few- femtosecond-duration soft x-ray pulses gener- ated through high-harmonic generation (24), has been successfully used to study ultrafast molecular relaxation processes with high structural and temporal resolution (25, 26). XTAS at the carbon K-edge is therefore an ideal platform to observe few-femtosecond timescale dynamics such as those associated + (27, 28). The x-ray with the JT distortion of CH4 probe excites C 1s electrons to unoccupied levels, such as the SOMO, or completely un- occupied antibonding or Rydberg levels. In particular, the dipole-allowed 1s → SOMO + is expected to be energetically signal in CH4 well resolved from other features. Geometric changes as a result of JT distortion will strongly affect the SOMO energy, which can be traced by XTAS with few-femtosecond time resolution. In this work, we report a joint experimental and theoretical study of the symmetry-breaking +. The cations were pro- JT dynamics of CH4 duced from neutral methane through abrupt, few-femtosecond SFI of CH4 with an 800-nm, few-cycle pump pulse generated by a table-top Ti:sapphire laser. The induced dynamics were then probed with XTAS using high-harmonic– generated soft x-ray pulses at the C K-edge obtained with a 1300-nm source (24). The non- perturbative nature of SFI leads to an ionization window that is temporally much narrower than the 5-fs width of the pump pulse (29). We observed a substantial energy shift in the XTAS signal immediately upon ionization because of JT distortion. The C2v minimum geometry was Ridente et al., Science 380, 713–717 (2023) 19 May 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. XTAS measurements of CH4 probe process. Pump-induced SFI produces Td symmetry CH4 JT distortion toward a C2v minimum. MOs for both structures are also shown (energies not to scale). The 1a1 orbital is the C 1s core-level, and the remainder are sC-H bonding orbitals. The dynamics are mapped by the 1s → SOMO probe + dynamics. (A) Schematic of the pump- +, which undergoes transition induced by a time-delayed (Dt) soft x-ray probe. (B) Ground-state carbon K-edge x-ray absorption spectrum for CH4. (C) Transient x-ray absorption at Dt < 35 fs. Negative time indicates probe preceding pump. The low-energy signal (278 to 282 eV) corresponds to the 1s → SOMO probe transition, which indicates that the C2v minimum is reached by ≈10 fs. attained within ≈10 fs, which was followed by two coherent oscillations in the signal that revealed large-amplitude scissoring motion. These coherent oscillations were damped out by ≈60 fs, indicating vibrational dephasing and eventual decoherence. The behavior of fully +) was also deuterated methane cations (CD4 investigated to understand the effect of sub- stituent masses on the JT dynamics. General features of XTAS signal CH4 has a simple ground-state x-ray absorp- tion spectrum (Fig. 1B). There is only one pro- minent peak (288 eV), which arises from the 1s → 3p Rydberg excitation (30). No other noteworthy pre-edge features were observed, and 1s ionization occurred at ≈290.8 eV (31). Figure 1C shows the experimental carbon + up to 35 fs K-edge XTAS spectrum for CH4 after the pump pulse abruptly ionizes the molecule at time t = 0. The intensity of the pump beam was 3 × 1014 W/cm2, which gener- + without substantial photo- ated sufficient CH4 dissociation, as discussed in the supplementary materials. Previous investigations have con- firmed that the major photoproduct for an 800-nm pump pulse of similar intensity and + (32). The negative (blue) pulse duration is CH4 transient signal (feature IV) corresponds to the depletion of neutral CH4 and can be more clearly observed at higher values of the pump power (see the supplementary materials). At t = 0, other prominent features were the po- sitive (red) signals at 278 to 282 eV (I), 284 eV (II), 287 eV (III), and 290 to 292 eV (V). The broad features II, III, and V had substantial temporal overlap with the pump pulse and could be attributed to the Stark effect of the pump pulse on core-excitation energies of CH4 (supplementary materials). Their tem- poral width is longer than the pump pulse because the Stark shift in XTAS measurements is proportional to the cross-correlation and timing jitter between pump and probe. After t = 0, a positive feature at ≈287.5 eV could be observed. This feature arose because of Raman activation of the symmetric stretch vibrational mode of CH4 by the pump pulse (supplemen- tary materials), like the behavior that has been observed in other molecules (33, 34). Feature I was assigned to CH4 + on the basis of orbital-optimized density functional theory (OO-DFT) calculations (35) that revealed that this feature corresponded to the 1s → SOMO +. OO-DFT excitation of nonequilibrium Td CH4 indicated that other core-level excitations of + (such as 1s → s* or the Rydberg levels) CH4 were above 288 eV in energy (supplementary materials). Feature VI in Fig. 1C corresponds to such excitations and could be observed at long times. However, feature VI did not show discernible time evolution. Conversely, feature I was well separated from all the other features and was particularly sensitive to changes in mo- lecular geometry as the SOMO is of sC-H char- acter. The time evolution of this signal was the +, and clearest reporter of the dynamics of CH4 the analysis below therefore focuses on it. + Relaxation dynamics of CH4 The long-time experimental XTAS of the JT feature corresponding to the 1s → SOMO tran- sition is given in Fig. 2A. The energetic average (henceforth abbreviated as CM1, for the first central moment) of the differential absorption [change in milli–optical density (DmOD)] signal (solid black line) showed three main charac- teristics. A rapid blue shift in energy from 278 eV (at t ≈ 0) to ≈282 eV (at t ≈ 18 fs) was followed by damped oscillations until t ≈ 60 fs and subsequently an almost time-independent sig- nal between 281 and 281.5 eV. The width of the spectral feature increased considerably starting around t ≈ 10 fs, leading to a very broad signal at longer times. We interpreted the behavior of this signal using quasiclassical ab initio molecular dyna- mics (AIMD) (36) trajectories on CH4 +. Figure 2B shows the XTAS signal computed using OO-DFT from 255 AIMD trajectory geometries at different time points, revealing good agree- ment with the experimental results. This comparison indicates that the trajectories under- lying the spectrum are a good reporter of the molecular dynamics under the experimental conditions. It is worth noting that all of the AIMD calculations were performed on the elec- + (i.e., were adia- tronic ground state of CH4 batic). Nonadiabatic effects from higher-energy + could potentially electronic states of CH4 contribute to the small differences between experiment and theory, but the experimental evidence suggests that the system is always in Ridente et al., Science 380, 713–717 (2023) 19 May 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E the ground electronic state under experimen- tal conditions (supplementary materials). In general, no explicit signature of nonadiaba- ticity was observed in the experimental spec- trum, as made evident by the good agreement between Fig. 2, A and B. The potential role of nonadiabatic effects on other aspects related to the JT distortion of CH4 has been considered elsewhere (9, 15–18). The timescale for the JT process can be di- rectly estimated by the time taken by the CM1 to attain the 281.5-eV value that we associate +. This value was with the C2v form of CH4 chosen because it corresponds to the asymp- totic long-time limit of the experimental signal CM1, and this value was also estimated by OO-DFT to be the upper bound for the 1s → + at the C2v equilibrium SOMO excitation of CH4 geometry (supplementary materials). This value was measured at 10 ± 2 fs from experiment by fitting an error function to the first 20 fs of the 1s → SOMO feature (supplementary mate- rials). OO-DFT calculations find this time to be 9.4 ± 0.3 fs, confirming that JT relaxation occurs on a timescale associated with high- frequency vibrational motion. In general, the time evolution of the signal corresponds to atomic motions associated with the relaxation process, with the oscillatory patterns suggesting involvement of vibrational +. This evolution could be anal- modes of CH4 yzed further using a Fourier transform (FT) of the CM1 position from both theory and experi- ment (Fig. 2C). Even though the FT features were broad owing to the rapid decay in the os- cillation amplitude, it was possible to identify critical frequencies. The most intense peak in the FT was at ≈1200 cm−1, a frequency that corresponded to a computed normal mode of the C2v minimum associated with scissoring about the H-C-H bond angles. This mode is an asymmetric scissoring mode (supplementary materials), where the scissoring motion about the smallest bond angle (i.e., the angle be- tween the two long C-H bonds) is opposite in direction to the scissoring motion about the largest bond angle (i.e., the angle between the two short C-H bonds). However, caution must be taken in interpreting the FT features in terms of the fundamental frequencies of the C2v + ground-state surface has minimum; the CH4 12 distinct C2v minima (37) and multiple seams corresponding to electronic state degeneracies, resulting in a highly anharmonic potential energy surface (PES). The FT is nonetheless an indication of several molecular motions that affect the signal and the associated timescales. The damping rate for the 1200 cm−1 frequency could also be estimated by fitting to the time domain experimental CM1 (Fig. 2D), revealing a lifetime of 58 ± 13 fs for the oscillations. We used the AIMD trajectories to uncover the origins of the signal oscillations in the x-ray spectra. The trajectories indicated that the C-H Fig. 2. Time evolution of the 1s → SOMO transition of CH4 +, with the first central moment (CM1) shown with the solid black line. The dotted black lines transition of CH4 indicate the asymptotic long-time signal CM1 (281.5 eV), which corresponds to the vibrationally hot C2v cation generated by the experiment, and at what time this energy is first reached (10 fs, JT timescale). (B) Theoretical XTAS for the same excitation. (C) FT of the CM1 from experiment (solid line) and theory (dashed line). The theoretical intensities have been uniformly scaled to match experiment for peak absorbance. (D) Experimental CM1 fit with −e−t/t × cos(wt), with w = 1200 cm−1, indicating a damping lifetime of t = 58 ± 13 fs for the vibrational dephasing. +. (A) Experimental XTAS for the 1s → SOMO Fig. 3. Role of the smallest H-C-H angle on XTAS signal. (A) Time evolution of the smallest bond angle over the trajectories. (B) Correlation between the computed 1s → SOMO excitation energies versus the smallest bond angle of the corresponding structures. (C) Evolution of the SOMO with change in the smallest bond angle. Ridente et al., Science 380, 713–717 (2023) 19 May 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E + and CH4 +. (A) Experimental Fig. 4. Comparison between CD4 XTAS of the 1s → SOMO transition of CD4 +, with the CM1 as the solid black line. The dotted lines indicate when the JT equilibrium energy is reached for the first time. (B) Theoretical XTAS for the same + from excitation. (C) Comparison of the CM1 for CH4 experiment (solid lines) and theory (dashed lines). The JT distortion is slower for CD4 lines indicate when the experimental signal CM1 first attained the value of 281.5 eV (corresponding to the C2v geometry) for CH4 + by a factor of ≈1.4. The vertical dotted +. + and CD4 + than CH4 + and CD4 bond lengths oscillate on a timescale roughly twice as fast as the principal oscillation in the CM1 (supplementary materials). The angular oscillations were slower than the stretches, with Fig. 3A showing that the mean of the smallest molecular bond angle over all trajec- tories underwent a damped oscillatory motion on the same timescale as the predicted and observed XTAS signal CM1 (Fig. 2, A and B). The computed bond angle distribution also broadened rapidly over time after t ≈ 10 fs, similar to the XTAS signal. A direct correla- tion between the observed signal and the smallest bond angle is revealed in Fig. 3B, which plots the smallest bond angle of the trajectory geometries used for Fig. 2B against the computed 1s → SOMO excitation energies for those geometries. It is evident that a simple linear model could capture much of the rela- tionship between the two quantities. The ex- citation energy had weaker correlation with other bond angles and bond lengths (supple- mentary materials). Theory therefore indi- cated that the most important contribution to the observed time evolution of the XTAS ab- sorption energy was from the dynamics of the smallest bond angle, to the extent that the signal could be interpreted in terms of a single parameter. This connection can be understood using a simple orbital model (Fig. 3C). The SOMO at + the C2v symmetry minimum geometry of CH4 is the bonding orbital arising from mixing be- tween a C 2p orbital and a symmetry-adapted linear combination (SALC) of the 1s orbitals corresponding to H atoms in the long bonds. This SALC has antibonding character because the two H 1s orbitals have opposite phases. For small bond angles, the H 1s SALC has poorer overlap with the C 2p level as it gets closer to the nodal plane of the latter. This behavior leads to a weaker interaction and therefore lowers the mixing between the C and H centered orbitals. Furthermore, smaller angles lead to decreased H-H distance, elevating the energy of the H 1s SALC because of the local anti- bonding character. The resulting sC-H bonding orbital therefore has greater nonbonding (pure C 2p) character as the angle decreases, made evident visually by the 30° angle case in Fig. 3C. Conversely, larger bond angles lead to a more stabilized SOMO with greater contribution from H orbitals. This picture is consistent with the observed increase in the 1s → SOMO x-ray probe excitation energy with decreasing bond angle shown in Fig. 3B. The x-ray oscillator strength also increased with a decrease in bond angle (supplementary materials). This behavior of the oscillator strength highlighted the increase in C 2p character of the SOMO because transitions from the C 1s level to other possible valence atomic orbitals that may contribute to the SOMO, such as C 2s or H 1s, have negligible oscillator strength. Therefore, the XTAS signal revealed the extent to which the SOMO lost C-H bonding character during the relaxation process. It is thus apparent that the Td → C2v JT distortion activates scissoring motion about the smallest bond angle in the C2v minimum, which was the most evident feature in the x-ray spectra. Furthermore, the frequencies indicated in Fig. 1C suggest that the asym- metric scissoring mode is activated to a greater extent than the symmetric scissoring mode. The greater activation of the asymmetric scis- soring mode is further supported by the short time (<10 fs) evolution of the molecular bond angles (supplementary materials), which revealed that the smallest H-C-H bond angle decreases over time, whereas the largest H-C-H bond angle increases in value. For a perfectly harmonic PES, the excess energy accumulated in this asym- metric scissoring mode would remain undissi- pated therein, leading to undamped oscillation of the XTAS signal and geometric parameters about the minimum before radiative relaxation to the vibrational ground state. However, the + is highly anharmonic (see discus- PES of CH4 sion above), and the surplus energy spreads out to all other modes. This redistribution was ob- served through damping in the oscillations for both the experimental XTAS signal CM1 and the mean geometric parameters from the AIMD trajectories (Figs. 2 and 3). In addition, consid- erable broadening of the XTAS signal was ob- served after the initial 10 fs, which was mirrored by an increase in the width of the probability distributions for the geometric parameters. Figure 2D indicates that the experimental XTAS CM1 oscillations had a damping lifetime of 58 fs, and oscillations in parameters computed from AIMD trajectories were mostly damped out within 60 fs (supplementary materials). It therefore appears that a large proportion of energy was transferred out of the JT-activated asymmetric scissoring mode to other internal degrees of freedom within this timescale, con- stituting an ultrafast example of intramolecular vibrational energy redistribution. This process can also be described in terms of dephasing of the asymmetric scissoring mode (i.e., relaxation Ridente et al., Science 380, 713–717 (2023) 19 May 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E toward thermal equilibrium), which was ulti- mately reflected in the decoherence of the XTAS signal. + and CD4 Comparison between experiment and theory revealed that the XTAS signal was reporting the nuclear motion over the duration of the re- laxation process. It is important to consider that SFI experiments could potentially exhibit electronic relaxation dynamics without consid- erable nuclear motion, involving highly excited Rydberg or cationic states (38, 39). To prove that the observed XTAS signal dynamics solely arose from nuclear (rather than electronic) relaxation, we performed experiments and computations for the dynamics after SFI of CD4. Figure 4A presents the time evolution of the resulting experimental XTAS signal, and Fig. 4B shows the computed XTAS spectrum from +. OO-DFT on 255 AIMD trajectories of CD4 The good agreement between experiment and theory further validated the ability of the com- puted trajectories to successfully simulate the experimental molecular dynamics. The greater mass of the deuterium isotope should slow the scissoring motion, leading to slower time evo- +. + relative to CH4 lution of the signal for CD4 Figure 4C confirms this to be the case by com- paring the time evolution of the CM1 for both + from experiment and theory. CH4 The time taken by the CM1 to reach the 281.5-eV value associated with the C2v cation geometry +, and the effect + than for CH4 was longer for CD4 of deuteration could be gauged by the ratio +) of these times. This ratio was (CD4 measured to be 1.4 ± 0.3 from experiment and calculated to be 1.44 ± 0.06 from theory (sup- plementary materials). The value thus calculated was close to the ratio between the asymmetric scissoring frequencies for the smallest bond + (computed to be 1.34). angle in CH4 However, it is important to note that all the + have similar fre- normal modes of C2v CH4 quency ratios upon isotopic substitution (sup- plementary materials). Nonetheless, the slower + unambiguously dynamics observed for CD4 reveal that the signal was reporting on nuclear dynamics. We also note that Fig. 4C shows that the relaxation dynamics after JT distor- + continue to be slower up to 45 fs, tion for CD4 from both experiment and theory. Longer time + XTAS signal is behavior for the computed CD4 + in the supplementary compared with CH4 materials (up to 85 fs), revealing slower coherent +. We also oscillations in the signal CM1 for CD4 note that Baker et al. (21) have reported that dynamics within the first 1.6 fs of ionization + versus were a factor of 2 to 3 slower for CD4 +. This behavior at very short times resulted CH4 from the decay of the autocorrelation of the nuclear wave function (9, 21), which is distinct from the longer time dynamics reported here involving substantial atomic displacements, therefore representing almost nonoverlapping nuclear wave functions. + and CD4 +/CH4 Conclusions + was prepared from SFI of CH4 and CH4 probed with XTAS near the carbon K-edge with few-femtosecond time resolution. Evo- lution of the excitation from the C 1s level to the valence hole revealed the dynamics of JT symmetry breaking away from the parent Td geometry, as well as subsequent coherent motion and dissipation of released energy out of active modes. All three of these aspects of intramolecular relaxation have been success- fully observed and analyzed. The combina- tion of experiment and theory revealed that the molecule first reached the JT-distorted form within 10 ± 2 fs after ionization. This distortion involved reduction of a H-C-H bond angle from 109.5° toward 55°, which was directly reported by a blue shift in the x-ray absorption signal. The JT dynamics were found to be 1.4× slower in deuterated methane on account of the larger substituent mass, proving that the observed dy- namics arose from nuclear motions. The energy released by the JT distortion drove a few coherent oscillations in the activated modes before being distributed over other molecular internal degrees of freedom, leading to damping of the oscillations within 60 fs of ionization. + We note that the observed behavior for CH4 is distinct from that observed in previous XTAS studies (26, 28) on the dynamics of CF4 + and + because those species are highly unstable CCl4 + has not been against bond dissociation. CF4 experimentally detected to date (26); meta- + has been previously observed (28), stable CCl4 but signals from the intramolecular relaxation pathways were unable to be disentangled from bond breaking. The subsequent coherence and dissipation of energy from the JT activated asymmetric scissoring mode to other internal +. degrees of freedom was only observed in CH4 This work thus opens the door to studies on how ultrafast vibrational coherence influences the redistribution of excess energy in more complex systems. RE FERENCES AND NOTES 1. H. A. Jahn, E. Teller, Proc. R. Soc. Lond. A 161, 220–235 (1937). I. B. Bersuker, Chem. Rev. 101, 1067–1114 (2001). 2. 3. H. C. Longuet-Higgins, in Advances in Spectroscopy, 4. H. W. Thompson, Ed. (Interscience Publishers, 1961), pp. 429–472. I. Persson, P. Persson, M. Sandström, A.-S. Ullström, Dalton Trans. 7, 1256–1265 (2002). 5. H. Köppel, L. S. Cederbaum, W. Domcke, J. Chem. Phys. 89, 2023–2040 (1988). 6. A. W. Potts, W. C. Price, Proc. R. Soc. Lond. A 326, 165–179 (1972). 7. L. B. Knight Jr., J. Steadman, D. Feller, E. R. Davidson, J. Am. Chem. Soc. 106, 3700–3701 (1984). 8. R. Signorell, M. Sommavilla, F. Merkt, Chem. Phys. Lett. 312, 139–148 (1999). 9. C. E. M. Gonçalves, R. D. Levine, F. Remacle, Phys. Chem. Chem. Phys. 23, 12051–12059 (2021). 10. E. Hirota, J. Mol. Spectrosc. 77, 213–221 (1979). 11. R. Signorell, F. Merkt, J. Chem. Phys. 110, 2309–2311 (1999). 12. R. Signorell, F. Merkt, Faraday Discuss. 115, 205–228 (2000). 13. H. J. Wörner, R. van der Veen, F. Merkt, Phys. Rev. Lett. 97, 173003 (2006). 14. H. J. Wörner, X. Qian, F. Merkt, J. Chem. Phys. 126, 144305 (2007). 15. T. Mondal, A. J. C. Varandas, J. Chem. Phys. 135, 174304 (2011). 16. T. Mondal, A. J. C. Varandas, J. Chem. Phys. 137, 214320 (2012). 17. P. Krause, S. Matsika, J. Chem. Phys. 136, 034110 (2012). 18. T. Mondal, A. J. C. Varandas, J. Chem. Phys. 143, 014304 (2015). 19. J. Rabalais, T. Bergmark, L. Werme, L. Karlsson, K. Siegbahn, Phys. Scr. 3, 13–18 (1971). 20. P. Baltzer et al., Chem. Phys. 196, 551–567 (1995). 21. S. Baker et al., Science 312, 424–427 (2006). 22. M. Li et al., Nat. Commun. 12, 4233 (2021). 23. L. Varvarezos et al., Chem. Phys. Lett. 775, 138687 (2021). 24. L. Barreau et al., Sci. Rep. 10, 5773 (2020). 25. R. Geneaux, H. J. B. Marroux, A. Guggenmos, D. M. Neumark, S. R. Leone, Phil. Trans. R. Soc. A 377, 20170463 (2019). 26. Y. Pertot et al., Science 355, 264–267 (2017). 27. K. S. Zinchenko et al., Science 371, 489–494 (2021). 28. A. D. Ross et al., Chem. Sci. 13, 9310–9320 (2022). 29. O. Smirnova, M. Spanner, M. Ivanov, J. Phys. B At. Mol. Opt. Phys. 39, S307 (2006). 30. J. Schirmer et al., Phys. Rev. A 47, 1136–1147 (1993). 31. J.-J. Pireaux et al., Phys. Rev. A Gen. Phys. 14, 2133–2145 (1976). 32. C. Wu et al., Phys. Rev. A 75, 043408 (2007). 33. A. Ferré et al., J. Phys. B At. Mol. Opt. Phys. 47, 124023 (2014). 34. K. F. Chang et al., J. Chem. Phys. 156, 114304 (2022). 35. D. Hait, M. Head-Gordon, J. Phys. Chem. Lett. 12, 4517–4529 (2021). 36. M. Paranjothy, R. Sun, Y. Zhuang, W. L. Hase, WIREs Comput. Mol. Sci. 3, 296–316 (2013). 37. M. N. Paddon-Row, D. J. Fox, J. A. Pople, K. N. Houk, D. W. Pratt, J. Am. Chem. Soc. 107, 7696–7700 (1985). 38. W. G. Scherzer, H. L. Selzle, E. W. Schlag, R. D. Levine, Phys. Rev. Lett. 72, 1435–1438 (1994). 39. Y. Kobayashi et al., Phys. Rev. A 101, 063414 (2020). 40. E. Ridente, Data for “Femtosecond Symmetry Breaking and Coherent Relaxation of Methane Cations via X-ray Spectroscopy,” data set, Zenodo (2023); https://doi.org/10. 5281/zenodo.7743367. AC KNOWLED GME NTS Funding: This work is funded by the US Department of Energy (DOE) Office of Science, Basic Energy Sciences (BES) Program, Chemical Sciences, Geosciences, and Biosciences Division, under contract no. DE-AC02-05CH11231, through the Gas Phase Chemical Physics program (E.R., E.A.H., A.D.R., D.M.N., and S.R.L.) and the Atomic, Molecular, and Optical Sciences program (D.H. and M.H.-G.). The instrument was built with funds from the National Science Foundation through NSF MRI 1624322 and matching funds from the Lawrence Berkeley National Laboratory, the College of Chemistry, the Department of Physics, and the Vice Chancellor for Research at UC Berkeley. This research used resources of the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility supported by the Office of Science of the US DOE under contract no. DE-AC02- 05CH11231 using NERSC award BES-ERCAP0020263. A.D.R. was additionally funded by the US DOE, Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering Division, under contract no. DE-AC02-05-CH11231, within the Physical Chemistry of Inorganic Nanostructures Program (KC3103); by the W. M. Keck Foundation, grant no. 042982; and by the US Army Research Office under grant no. W911NF-20-1-0127. Author contributions: Experimental investigation: E.R., E.A.H., and A.D.R. Theoretical investigation: D.H. Experiment supervision: S.R.L. and D.M.N. Theory supervision: M.H.-G. Writing – original draft: E.R., D.H., and E.A.H. Writing – review & editing: E.R., D.H., E.A.H., A.D.R., M.H.-G., D.M.N., and S.R.L. Competing interests: The electronic structure calculations were performed in Q-Chem, which is partially owned by M.H.-G. The other authors declare no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper or the supplementary materials. The experimental and theoretical data can be accessed on Zenodo (40). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg4421 Materials and Methods Figs. S1 to S25 Tables S1 to S3 References (41–64) Submitted 23 December 2022; accepted 19 April 2023 10.1126/science.adg4421 Ridente et al., Science 380, 713–717 (2023) 19 May 2023 5 of 5
10.1126_science.adf7122
RES EARCH BIOGEOGRAPHY Paleoenvironments shaped the exchange of terrestrial vertebrates across Wallace’s Line A. Skeels1,2,3*, L. M. Boschman4, I. R. McFadden1,2,5, E. M. Joyce6, O. Hagen7, O. Jiménez Robles3,8, W. Bach1,2, V. Boussange1,2, T. Keggin1,2, W. Jetz9,10, L. Pellissier1,2 Faunal turnover in Indo-Australia across Wallace’s Line is one of the most recognizable patterns in biogeography and has catalyzed debate about the role of evolutionary and geoclimatic history in biotic interchanges. Here, analysis of more than 20,000 vertebrate species with a model of geoclimate and biological diversification shows that broad precipitation tolerance and dispersal ability were key for exchange across the deep-time precipitation gradient spanning the region. Sundanian (Southeast Asian) lineages evolved in a climate similar to the humid “stepping stones” of Wallacea, facilitating colonization of the Sahulian (Australian) continental shelf. By contrast, Sahulian lineages predominantly evolved in drier conditions, hampering establishment in Sunda and shaping faunal distinctiveness. We demonstrate how the history of adaptation to past environmental conditions shapes asymmetrical colonization and global biogeographic structure. of environmental gradients between regions (9, 13). To date, our ability to distinguish be- tween hypotheses regarding the processes that shape the unevenness of different biotic interchanges has been limited by the avail- able process-based methodologies and paleo- environmental reconstructions. Distributional patterns of terrestrial verte- brates in the Indo-Australian archipelago have fascinated naturalist for centuries (14) and have provided a model system for understand- ing biotic interchange. As the Australian con- tinental plate approached the Eurasian plate, subduction at the northern boundary led to the formation of a geologically complex archi- pelago, today composed of thousands of islands and known biogeographically as Wallacea (5) (Fig. 1). The oceanic boundary separating the Sunda continental shelf (including Myanmar, Cambodia, Vietnam, Laos, Thailand, Malaysia, and Indonesia west of Lombok on the Eurasian plate) from Wallacea and the Sahul continental shelf (including Australia and New Guinea on the Australian plate) is named Wallace’s Line after Alfred Russell Wallace (14) (Fig. 1), who noted a clear disjunction in the distribution of some taxa across it (14, 15). Today, Wallace’s Line marks the boundary of the Indomalayan and Australasian zoogeographic realms (2, 16) (Fig. 1), but there are many lines in the region that mark turnover between a predominantly Sundanian or Sahulian fauna (15), such as the Heilprin-Lydekker Line (Fig. 1). The persistent presence of oceanic barriers has led to the dominant idea that dispersal limitation is the primary process shaping interchange (12). For example, the mammalian faunas of Sunda and Sahul are largely endemic, and only vagile rodents (17) and bats (18) have been successfully exchanged between continents. However, envi- ronmental niche traits may also have played a role (19, 20) as the distinctly drier Australian continent (20, 21) shaped the evolution of T he world’s major biotic interchanges have had a disproportionate impact on the dis- tribution of the world’s terrestrial fauna (1–3). Well-known systems include the American interchange, which followed the formation of the Panamanian isthmus be- tween North and South America (4), and the Indo-Australian interchange, which followed the convergence of the Australian and Eur- asian tectonic plates and was shaped by com- plex dynamics of island formation (5). Far from an even diffusion of species between re- gions, interchanges have displayed an asym- metry in the direction of exchange (6–8) and have been dominated by particular functional traits (9, 10). This unevenness may reflect a predisposition for dispersion in particular groups of organisms, because successful colo- nization is dependent on the ability both to disperse to a new region and to establish within new environmental conditions and ecological communities (11). Dispersal ability may act as a primary filter on colonization success (12), but lineages with broad environ- mental tolerances, or those with the ability to adapt to new environments, might contrib- ute more to an interchange in the presence 1Department of Environmental Systems Science, Ecosystems and Landscape Evolution, Institute of Terrestrial Ecosystems, ETH Zurich, 8092 Zurich, Switzerland. 2Swiss Federal Institute for Forest, Snow and Landscape Research WSL, 8903 Birmensdorf, Switzerland. 3Research School of Biology, Australian National University, Canberra 0200, Australia. 4Department of Earth Sciences, Utrecht University, 3584 CB Utrecht, Netherlands. 5Institute for Biodiversity and Ecosystem Dynamics, University of Amsterdam, 1090 GE Amsterdam, Netherlands. 6Systematics, Biodiversity and Evolution of Plants, Ludwig Maximilian University of Munich, 80331 Munich, Germany. 7German Centre for Integrative Biodiversity Research (iDiv) Halle-Jena-Leipzig, 04103 Leipzig, Germany. 8Institute of Biology, École Normale Supérieure, 75005 Paris, France. 9Department of Ecology and Evolutionary Biology, Yale University, New Haven, CT 06520, USA. 10Center for Biodiversity and Global Change, Yale University, New Haven, CT 06520, USA. *Corresponding author. Email: alexander.skeels@gmail.com Skeels et al., Science 381, 86–92 (2023) 7 July 2023 Wallace’s Line Heilprin-Lydekker Line Arid and temperate Tropical rainforest Tropical savanna Sunda Wallacea New Guinea Indomalayan Realm Australasian Realm Sahul Australia 120°E 140°E 0° 10°S 20°S Fig. 1. Climate and geography of the Indo-Australian archipelago at present. The Sunda continental shelf is part of the Indomalayan zoogeographic realm and is separated from the Australasian zoogeographical realm by Wallace’s Line. Wallacea is the island archipelago between the Sunda and Sahul continental shelves. The Sahul continental shelf, which includes the Australian mainland and New Guinea, is separated from Wallacea by the Heilprin-Lydekker Line. Sunda, Wallacea, and New Guinea are dominated by the tropical rainforest Köppen-Geiger climate belt, except for eastern Java, Bali, South Sulawesi, the Lesser Sunda isles, and southern New Guinea, which are dominated by the drier tropical savanna climate belt. The Australian mainland is dominated by tropical savanna, arid, and temperate climate belts. The climate maps are adapted from (48). 1 of 6 RES EARCH | R E S E A R C H A R T I C L E an arid-adapted fauna (22). The interchange has been highly asymmetrical, with rates of exchange estimated to be at least twice as high from Sunda to Sahul than in the oppo- site direction (23). This could be shaped by the uneven sizes of the species pools, as stable areas of tropical rainforest in Sunda, connected to the Eurasian continent, supported a bio- diverse fauna throughout the Cenozoic (24). Conversely, increasing aridity and a less connected continental area bound by oceans may have led to a reduced species pool in Sahul. Our current understanding of interchanges is pat- tern oriented, with historical biogeographic analyses determining where, when, and in which direction exchanges occur (2, 12, 15, 23, 25). A process-oriented understanding, however, requires the synthesis of paleogeographic and paleoclimatic information with biological traits and diversification histories across a range of taxa. Paleoenvironmental and biological informa- tion united under a mechanistic modeling approach makes it possible to compare empir- ical observations with theoretical expectations and provides a framework with which to ex- plore the mechanisms behind one of the world’s major biotic interchanges. We hypoth- esize that, beyond a simple dispersal filter across oceanic barriers, paleoenvironmental niche dynamics also influence patterns of ex- change. In addition to each clade’s dispersal ability (12) and the size of the species pool, exchange patterns should be related to either the variability in the environmental conditions in which species occur (25) (mean realized niche breadth) or the rate at which environ- mental niche traits change over time (13) (rates of niche evolution). Further, the spa- tial distribution of climate, and how it has changed through time, should also explain asymmetry and the spatial distribution of lineages with either Sundanian or Sahulian ancestry (19). Here, we combined paleoenvi- ronmental reconstructions of temperature, precipitation, and plate tectonics over the past 30 million years (Ma) (26–28) with a mecha- nistic model of biodiversity (29) to explore how dispersal and niche traits interact with the environment to shape regional diversifi- cation and patterns of exchange through time (30). The selected time period captures the on- set of the convergence of the Australian and Eurasian plates (5) and the origin of most ter- restrial vertebrate families in the region (fig. S1). We used published, well-sampled phylogenetic and spatial information on 20,433 species be- longing to all 227 families of terrestrial vertebrates present in the Indo-Australian archipelago to estimate phylogenetic and taxonomic b di- versity across Wallace’s Line as measures of compositional dissimilarity between the two regions, as well as the number and direction of colonization events (31). Trait-based drivers of faunal exchange We estimated a minimum of 381 independent colonization events across Wallace’s Line based on a conservative phylogenetic hypothesis for each family [figs. S2 to S5 (30)], explaining the widespread distribution across Sunda and Sahul of 87 out of 227 terrestrial vertebrate families (Fig. 2 and Fig. 3A). Older colonization events were associated with greater species diversity (Spearman r = 0.68; Fig. 2A), most likely because these older lineages had a longer time to diversify. Precipitation niche breadth was the most consistent predictor of exchange in the Indo-Australian archipelago, indicating that colonization is more common in lineages that can tolerate wider variation in precipitation. Using phylogenetic multiple regression (table S1), we found that families with broader pre- cipitation niche breadths were more likely to be exchanged between regions, to have a larger number of exchange events, and to have lower taxonomic b diversity and phyloge- netic b diversity (Fig. 3B and table S1) across Wallace’s Line. Temperature niche breadth was negatively associated with exchange dy- namics (Fig. 3B and table S1), because tem- perature and precipitation niche breadths are inversely related in vertebrates (32). The 20 Region of origin a M / Sunda Sahul A B i ) y t i s r e v d s e c e p s ( n i l 10 0 20 10 a M / s t n e v e n o i t i a z n o o C l C Line Heilprin-Lydekker Wallace d e s s o r c s e i l i m a f . N 100 50 0 Sunda Sahul Origin Oligocene 30 Miocene Pliocene Pleistocene 0 10 0 20 Age (Ma) ClimPC1 10 N 150 0 100 10 S 50 20 S 30.0 Ma 22.5 Ma 15.0 Ma 7.5 Ma 0.0 Ma 90 E 100 E 110 E 120 E 130 E 140 E 150 E 90 E 100 E 110 E 120 E 130 E 140 E 150 E 90 E 100 E 110 E 120 E 130 E 140 E 150 E 90 E 100 E 110 E 120 E 130 E 140 E 150 E 90 E 100 E 110 E 120 E 130 E 140 E 150 E Fig. 2. Colonization dynamics of terrestrial vertebrates and geoclimatic changes in Indo-Australia. (A) Bars show the log-transformed number of species derived from independent colonization events across Wallace’s Line from Sunda to Wallacea (blue) or in the opposite direction (yellow). Dashed lines show the number of colonization events through time. (B) Paleogeography (28) and paleoclimate (26, 27) of the Indo-Australian archipelago over the past 30 Ma (26, 28). (C) Numbers of families crossing between regions across two different biogeographic boundaries. Principle component analysis was performed on paleotemperature and paleoprecipitation to decompose a single climate axis (ClimPC1), explaining 88% of the variation in climate, where similar colors represent similar climates. Warm, wet conditions are shown in red; cold, dry conditions are shown in blue. Skeels et al., Science 381, 86–92 (2023) 7 July 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E mechanistic model enabled the disentangle- ment of these factors and validated the role of broader precipitation niche breadths in the rate of exchange (table S2). Niche breadth may affect colonization dynamics through two pathways: (i) Clades more tolerant of a broad range of conditions may be less prone to ex- tinction and therefore have a larger number of potential dispersing lineages (8), or (ii) they may be more able to establish in a diverse range of environmental conditions with greater establishment success. Our mechanistic model supports the second pathway, because higher rates of colonization were not associated with lower rates of extinction (fig. S6), but instead associated with certain dispersal and niche characteristics of the modeled lineages. Al- though the simulation results cannot exclude the role of differential extinction on coloni- zation dynamics [e.g., (8)], they do suggest that differences in the rate of lineage diversifica- tion are not required to explain asymmetrical exchange. These results support the hypoth- esis that the precipitation gradient between Sunda and Sahul is a sharp barrier to the estab- lishment of range-expanding lineages (10, 33), and that lineages able to persist in a diversity of conditions can overcome this barrier. Supplementing the effect of climatic niche breadth, our empirical and in silico model results support the importance of dispersal ability and the species pool in exchange pat- terns. More diverse biotas might facilitate in- terchange by having a greater number and variation of potential colonist species, increas- ing the number of potential colonization at- tempts (34), and this pathway is considered the null expectation in invasion biology (35). We found that the size of the source species pool was positively associated with exchange success and the number of colonization events in terrestrial vertebrates (Fig. 3B and table S1), as well as with all exchange metrics in the mechanistic model (table S2). As a key dis- persal trait, flight was negatively associated with phylogenetic b diversity (Fig. 3B), and most exchanged taxa were highly volant, with 64% of families being either birds or bats, A Prec. breadth (mm/yr) 1000 500 Origin 48 Sahul Sunda Equivocal Exchanged? Yes No 47 46 45 44 43 42 41 40 39 38 37 36 35 34 33 32 31 30 3 2 4 5 1 6 7 8 9 10 11 12 B Succesful exchange 13 14 15 16 17 18 19 20 -1 Number of colonizations -2 Taxonomic ß−diversity 0 0 1 2 350 Time (Ma) 250 150 50 -0.05 0.00 0.05 Phylogenetic ß−diversity 29 8 2 7 2 6 2 5 2 4 2 3 2 2 2 21 -0.10 -0.05 0.00 Coefficient estimate 0.05 Temperature rate Temperature breadth Precipitation rate Precipitation breadth Species pool Dispersal ability Temperature rate Temperature breadth Precipitation rate Precipitation breadth Species pool Dispersal ability Temperature rate Temperature breadth Precipitation rate Precipitation breadth Species pool Dispersal ability Temperature rate Temperature breadth Precipitation rate Precipitation breadth Species pool Dispersal ability 0.10 Fig. 3. Exchange dynamics and precipitation niche breadth of terrestrial vertebrates on a family-level phylogenetic tree and effect size of the predictors of exchange. (A) The circle in the center displays each family’s reconstructed origin as Sunda, Sahul, or uncertain (equivocal support) based on biogeographic estimation models. The middle colored ring shows whether the family is present in both Sunda and Sahul and was reconstructed to have been exchanged across Wallace’s Line. In the outer colored ring, the color and height of the bars represent the mean precipitation niche breadth of the family. Niche breadth was measured as the SD of mean annual rainfall across each species distribution (in millimeters per year). The numbered lines at the perimeter indicate the number of taxonomic orders, with representative taxa highlighted as silhouettes (images are from PhyloPic: http://phylopic.org/). 1, Carnivora; 2, Pholidota; 3, Cetartiodactyla; 4, Perrissodactyla; 5, Chiroptera; 6, Eulipotyphla; 7, Primates; 8, Scandentia; 9, Dermoptera; 10, Rodentia; 11, Lagomorpha; 12, Probosoidea; 13, Diprotodontia; 14, Dasyuromorphia; 15, Peramelemorphia; 16, Notoryctemorphia; 17, Monotremata; 18, Anura; 19, Caudata; 20, Gymnophiona; 21, Passeriformes; 22, Psittaciformes; 23, Falconiformes; 24, Piciformes; 25, Coraciiformes; 26, Bucerotiformes; 27, Trogoniformes; 28, Strigiformes; 29, Accipitriformes; 30, Charadriiformes; 31, Apodiformes; 32, Caprimulgiformes; 33, Eurypygiformes; 34, Suliformes; 35, Pelecaniformes; 36, Ciconiiformes; 37, Procellariiformes; 38, Sphenisciformes; 39, Gruiformes; 40, Cuculiformes; 41, Otidiformes; 42, Phaethontiformes; 43, Podicipediformes; 44, Columbiformes; 45, Galliformes; 46, Anseriformes; 47, Casuariiformes; 48, Squamata. (B) Coefficients of phylogenetic linear models of four exchange metrics with mean temperature and precipitation niche breadths, rates of temperature and precipitation niche evolution, dispersal ability, and species pool size. Predictors that were statistically significant based on phylogenetic linear models are in bold, and nonsignificant predictors are gray. Skeels et al., Science 381, 86–92 (2023) 7 July 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E A ) C ° ( e r u t a r e p m e T B ) C ° ( e r u t a r e p m e T C ) C ° ( e r u t a r e p m e T 25 20 15 10 5 0 25 20 15 10 5 0 25 20 15 10 5 0 Sunda Niche centroids exchanged non-exchanged Wallacea Age (Ma) 30 20 10 0 5 10 15 Prepicitation (mm/day) Sahul Wallacea Age (Ma) 30 20 10 0 5 10 15 Prepicitation (mm/day) Wallacea Proportion Sahulian 1.00 0.75 0.50 0.25 0.00 0 5 10 15 Prepicitation (mm/day) Fig. 4. Age of paleoclimatic space in the Indo- Australian archipelago and distribution of climatic niche centroids of terrestrial vertebrates. (A and B) Gridded representation of climate space defined by temperature and precipitation. Each tile in the grid represents a unique combination of these two variables. The tiles are colored by the age of that climate’s first appearance in Sunda (A) and Sahul (B). Climate spaces that have been present longer are in darker colors. The climatic space occupied by the “stepping stone” region of Wallacea during the past 30 Ma is highlighted with an orange dashed line. This space has been present longer in Sunda than in Sahul. Symbols indicate the median climate niche centroid of species from Sahul (yellow) or Sunda (blue) that belong to families that have been exchanged (circle) or have not been exchanged (star). (C) Proportion of species of Sahulian origin in climate space. Species niche positions in climate space show that the area occupied by Wallacea is more densely filled with species from Sundanian families than with species from Sahulian families, and that species from both exchanged and nonexchanged Sundanian families are more commonly found in the climate space occupied by Wallacea. Skeels et al., Science 381, 86–92 (2023) 7 July 2023 accounting for 93% of the colonization events of Sunda. All exchange metrics were strongly positively associated with dispersal ability in the mechanistic model (table S2). We found that exchange of bird families across the Indo- Australian archipelago was proportionally higher than during other major biotic inter- changes (36), reflecting the greater propensity of volant taxa to disperse over the deep-sea oceanic barriers such as Wallace’s Line, which uniquely define the Indo-Australian interchange. Contrary to our prediction, rates of niche evolution were not commonly associated with exchange dynamics (Fig. 3B and tables S1 and S2), possibly reflecting the widespread pheno- menon of niche conservatism, the tendency for lineages to retain their ancestral environ- mental niche through time (37), during trans- oceanic colonization (38). For instance, many families that we reconstructed as having a Sundanian origin, such as Colubrid snakes and Pteropodid bats, are typically restricted to similar biomes in Sahul. Instead of the ability to rapidly adapt to new environmental condi- tions, we show that broader precipitation niche breadths and higher dispersal ability were the primary biological traits that allowed line- ages to expand their geographic ranges in the face of oceanic barriers and a steep precip- itation gradient. Paleoclimate and exchange asymmetry We demonstrate that asymmetrical rates of exchange have been shaped by the distinctly different paleoclimatic histories of Sunda and Sahul. At least 46 families originating in Sunda (43% of all Sundanian families) crossed the Heilprin-Lydekker Line (Fig. 1 and table S3) to Sahul, whereas only 19 families originating in Sahul (24% of all Sahulian families) crossed Wallace’s Line to Sunda. A necessary step to colonization involves crossing Wallacea (Fig. 1). Our paleoenvironmental reconstruction, which combines paleotopography (28), reconstructed Köppen belts (27), and a global circulation model (26), show that the humid climate space occupied by Wallacea emerged in Sunda at least 20 and 30 Ma (Fig. 4A and fig. S7) in a period associated with the spread of tropical rainforests (21, 24). Conversely, in Sahul, areas with similarly humid conditions were present during the mid-Miocene (fig. S8) (39) and only became widespread in their contemporary lo- cation on Sahul at least 10 and 5 Ma (Fig. 4B), after the northward movement of Sahul and the uplift of New Guinea (20, 24). An almost uninterrupted humid climate throughout the Indo-Australian archipelago, from Sunda to New Guinea, facilitated the establishment of Sundanian lineages in Sahul, and this can also explain the observation that Wallacean and New Guinean flora are compositionally more similar to the flora of Sunda than to that of the Australian mainland (33). Support- ing this finding, we show that the climate space currently present in Wallacea and New Guinea is more densely occupied by vertebrate lineages from Sunda than by lineages from Sahul (Fig. 4C). By contrast, species from Sahul in families that did not colonize Sunda occupy an older and drier component of the Sahulian climate than species in families that colonized Sunda (Fig. 4C and fig. S9). A higher density of species in older climate space suggests that niche evolution has been truncated through time by the available climate on each continent, highlighting that niche conservatism limits exchange between Sunda and Sahul. Greater connectivity of a humid climate at the interface of the exchange explains the ob- served patterns of asymmetrical interchange in terrestrial vertebrates, and this was sup- ported by the mechanistic model. In the mod- el, evolution of climatic niches was directed by the environmental conditions where species evolved (fig. S7). Therefore, Sundanian line- ages evolved in niches closer to the condi- tions of Wallacea and colonized emergent Wallacean islands earlier (mean colonization = 15.27 Ma; fig. S10). In the opposite direction, Sahul lineages were generally unable to suc- cessfully colonize proto-Wallacea. Coloniza- tion events by Sahulian species across Wallace’s Line generally occurred after the uplift of New Guinea’s highlands (mean colonization time = 5.4 Ma; fig. S10), supporting the role of New Guinea as an ecological “stepping stone” between Wallacea and the Australian mainland (40). We evaluated how climate connectivity ex- plains exchange asymmetry by performing an in-silico counter-factual simulation experi- ment. Here, we either removed the precipi- tation or temperature constraints on species distributions or removed the tectonic history by holding the landscape constant with sea levels at the time period with maximum con- nectivity between the regions (Last Glacial Maximum, ~20,000 years ago). This effectively allowed dispersal within each continental re- gion to be unconstrained by oceanic barriers and provided equivalent geographic isolation of each continental region from Wallacea. We found that the proportion of Sahulian species in Wallacea and New Guinea increased only when both the precipitation constraint and tectonic history were modified together, be- cause this promoted more opportunities for dispersal and a higher likelihood of establish- ment (Fig. 5 and figs. S11 and S12). The propor- tion of simulations with colonization events between Sahul and Sunda was 16.6% in the full model, which was similar to the observed proportion in terrestrial vertebrates (24%). The proportion increased to 48.6% in the counterfactual model (fig. S11), which is more similar to the observed proportion of line- ages moving in the opposite direction, from 4 of 6 RES EARCH | R E S E A R C H A R T I C L E A Terrestrial vertebrates B Birds 20 N 10 N 0 10 S 20 S 30 S 40 S Prop. of species from Sahulian families 1.00 0.75 0.50 0.25 0.00 M5 M6 C Amphibians M3 M1L M1 M4 M2 M1H D Mammals E Reptiles 80 E 90 E 100 E 110 E 120 E 130 E 140 E 150 E Fig. 5. Spatial distribution of the proportion of terrestrial vertebrate species of Sahul origin and biogeographic turnover predictions from the biodiversity simulation model. Proportion of total species richness in each grid cell that belongs to families of Sahul origin, based on biogeographic estimation, over the count of all species in the grid cell, for all terrestrial vertebrates (A), birds (B), amphibians (C), mammals (D), and reptiles (E). Dashed lines in (A) show the location of the sharpest spatial turnover of proportional Sahulian richness under six different simulations models (M1 to M6), as well as under high (M1H) and low (M1L) values of precipitation niche breadth and dispersal ability parameters of the full model (M1). The precipitation constraint on species distributions is removed in M2, the temperature constraint is removed in M3, tectonic history is removed in M4, both the precipitation constraint and tectonic history are removed in M5, and both the temperature constraint and tectonic history are removed in M6. The dashed lines illustrate how removing certain constraints that limit colonization changes the distribution of species from families with either Sahulian or Sundanian ancestry. Grid cell size is 110 × 110 km for all panels (Behrmann projection). Sunda to Sahul (43%). Therefore, more sym- metrical exchange could only be observed in the absence of climatic and dispersal bar- riers. A similar climate-mediated scenario has been suggested to explain exchange asym- metry during the American interchange (9, 41) and between India and Eurasia (7, 42). Our study clarifies how distinct climatic histories of continental regions can generate a near- ubiquitous pattern of asymmetrical terres- trial biotic interchanges. Dispersal and environmental niche shape spatial patterns of faunal exchange Patterns of turnover among taxa have long been a source of contention for designating bio- geographic boundaries in the Indo-Australian archipelago (15) because there is a high dis- cordance in the distributional extent of differ- ent taxa across the Indo-Australian archipelago. Across the four classes of terrestrial vertebrates, gradients of turnover from a predominantly Sundanian to a predominantly Sahulian fauna differed, with a high proportion of sites in Sahul having high richness of Sundanian rep- tile species (100% of sites; Fig. 5E) and mam- mal species (68.8%; Fig. 5D) (which is driven by the exceptional diversity of Sundanian squamates, rodents, and bats), but not for amphibian species (11%; Fig. 5C) or bird spe- cies (1.4%; Fig. 5B). In the mechanistic model, a combination of dispersal and niche traits shaped gradients in biotic turnover, which could explain the emergence of these differ- ent patterns in vertebrates. When simulations were run with increased species’ precipitation niche breadth and dispersal ability parame- ters, we found that lines demarcating turn- over from a Sundanian to Sahulian biota (Fig. 5A) were located southward into regions of increased aridity, varying from a pattern corre- sponding with amphibians (Spearman’s r = 0.85 to 0.87; Fig. 5C) to a pattern correspond- ing with mammals (r = 0.71 to 0.74; Fig. 5D), respectively. For birds, New Guinea is predom- inantly Sahulian, and this pattern was only reproduced by completely lifting the dispersal and precipitation constraints from the model Skeels et al., Science 381, 86–92 (2023) 7 July 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E (r = 0.96; Fig. 5B and fig. S11). This could be because birds have the broadest precipitation tolerances and the highest volancy of any class of vertebrate, and therefore have fewer con- straints on colonization (fig. S13). Because the simulation model is agnostic to the taxa in- vestigated, we expect a similar mechanism to operate across a range of taxa not explored here, and there is some evidence that niche and dispersal traits are important drivers of interchange and patterns of turnover in plants (10, 43). Conclusion The longstanding view of biotic interchange is that it is primarily governed by plate tec- tonics and dispersal ability, with rates of colo- nization in terrestrial organisms being directly proportional to the source pool size (1), the geographic distance between regions, and the emergence of land bridges (9). A more com- plete picture emerges when the deep-time legacy of climate connectivity and niche con- servatism on species distributions are also considered. Advances in deep-time climate re- constructions make it possible to test this directly, and support for a key role of climate connectivity is emerging in the Indo-Australian archipelago (33) and more broadly in other systems such as Indo-Eurasia (7, 42) and the American interchange (41). The factors that shape colonization success are relevant in both a historical and a contemporary context, because biotic exchange is exacerbated by human intervention, and understanding these factors is therefore relevant for predicting the success of new biological colonizations and invasions (35). This is particularly true in the Indo-Australian archipelago, where bio- logical ivasions are contributing to some of the highest rates of native species extinction on the planet (45). An emerging view of global biodiversity patterns is that they are largely the outcome of historical changes in plate tectonics shuffling lineages around the planet, in concert with major fluctuations in precip- itation and temperature shaping dispersal, speciation, and extinction dynamics through environmental niches (46, 47). A mechanistic modeling approach enables us to move be- yond “lines” in biogeography and instead con- sider the processes that shape global variation in biodiversity patterns. RE FERENCES AND NOTES 1. G. J. Vermeij, Science 253, 1099–1104 (1991). 2. B. G. Holt et al., Science 339, 74–78 (2013). 3. G. F. Ficetola, Nat. Ecol. Evol. 1, 89 (2017). 4. F. G. Stehli, S. D. Webb, The Great American Biotic Interchange (Springer, 2013), vol. 4. 5. D. J. Lohman et al., Annu. Rev. Ecol. Evol. Syst. 42, 205–226 (2011). 6. S. D. Webb, Paleobiology 17, 266–280 (1991). 7. S. Klaus, R. J. Morley, M. Plath, Y. P. Zhang, J. T. Li, Nat. Commun. 8. 9. 7, 12132 (2016). J. D. Carrillo et al., Proc. Natl. Acad. Sci. U. S. A. 117, 26281–26287 (2020). J. T. Weir, E. Bermingham, D. Schluter, Proc. Natl. Acad. Sci. U. S. A. 106, 21737–21742 (2009). 10. J. Y. S. Yap et al., J. Biogeogr. 45, 838–847 (2018). 11. J. P. Sexton, P. J. McIntyre, A. L. Angert, K. J. Rice, Annu. Rev. Ecol. Evol. Syst. 40, 415–436 (2009). 12. A. E. White, K. K. Dey, M. Stephens, T. D. Price, Glob. Ecol. Biogeogr. 30, 685–696 (2021). 13. W. L. Eiserhardt, J. C. Svenning, W. J. Baker, T. L. P. Couvreur, H. Balslev, Sci. Rep. 3, 1164 (2013). 14. A. R. Wallace, Proc. R. Geogr. Soc. Lond. 7, 206–213 (1863). 15. J. R. Ali, L. R. Heaney, Biol. Rev. Camb. Philos. Soc. 96, 922–942 (2021). 16. D. M. Olson et al., Bioscience 51, 933–938 (2001). 17. K. C. Rowe et al., J. Biogeogr. 46, 2752–2768 (2019). 18. S. M. Tsang et al., J. Biogeogr. 47, 527–537 (2020). 19. R. Good, Nature 181, 1763–1765 (1958). 20. H. A. Nix, in Evolution of the Flora and Fauna of Arid Australia, W. R. Barker, P. J. M. Greenslade, Eds. (Australian Systematic Botany Society, 1982), pp. 47–66. 21. R. J. Morley, J. Trop. Ecol. 34, 209–234 (2018). 22. M. Byrne et al., Mol. Ecol. 17, 4398–4417 (2008). 23. D. M. Crayn, C. Costion, M. G. Harrington, J. Biogeogr. 42, 11–24 (2015). 24. R. Morley, in Tropical Rainforest Responses to Climatic Changes, M. Bush, J. Flenley, W. Gosling, Eds. (Springer, 2011), pp. 1–34. 25. K. A. Carscadden et al., Q. Rev. Biol. 95, 179–214 (2020). 26. P. J. Valdes, C. R. Scotese, D. J. Lunt, Clim. Past 17, 1483–1506 (2021). 27. C. R. Scotese, H. Song, B. J. W. W. Mills, D. G. van der Meer, Earth Sci. Rev. 215, 103503 (2021). 28. R. Hall, J. Limnol. 72 (s2), e1 (2013). 29. O. Hagen et al., PLOS Biol. 19, e3001340 (2021). 30. Materials and methods are available as supplementary materials. 31. N. J. Matzke, Front. Biogeogr. 5, 242–248 (2013). 32. M. R. Pie, R. Divieso, F. S. Caron, J. Zool. Syst. Evol. Res. 59, 1155–1166 (2021). 34. J. L. Lockwood, P. Cassey, T. Blackburn, Trends Ecol. Evol. 20, 223–228 (2005). 35. G. J. Vermeij, Biol. Conserv. 78, 3–9 (1996). 36. C. D. Bacon et al., Proc. Natl. Acad. Sci. U.S.A. 112, 6110–6115 (2015). 37. J. J. Wiens, Evolution 58, 193–197 (2004). 38. M. D. Crisp et al., Nature 458, 754–756 (2009). 39. K. J. Travouillon, S. Legendre, M. Archer, S. J. Hand, Palaeogeogr. Palaeoclimatol. Palaeoecol. 276, 24–37 (2009). 40. E. Roycroft et al., Curr. Biol. 32, 4215–4224.e3 (2022). 41. C. D. Bacon et al., Geology 44, 375–378 (2016). 42. M. Bansal et al., Science 375, 455–460 (2022). 43. L. A. Trethowan, J. Veg. Sci. 32, 1–7 (2021). 44. S. C. Bernardes et al., Sci. Rep. 11, 23357 (2021). 45. J. C. Z. Woinarski, A. A. Burbidge, P. L. Harrison, Proc. Natl. Acad. Sci. U.S.A. 112, 4531–4540 (2015). 46. O. Hagen, A. Skeels, R. E. Onstein, W. Jetz, L. Pellissier, Proc. Natl. Acad. Sci. U.S.A. 118, e2026347118 (2021). 47. O. Hagen, Ecography 4, e06132 (2022). 48. H. E. Beck et al., Sci. Data 5, 180214 (2018). 49. A. Skeels et al., Data for: Paleoenvironments shaped the exchange of terrestrial vertebrates across Wallace’s Line, Dryad (2023); https://doi.org/10.5061/dryad.1c59zw41f. AC KNOWLED GME NTS We thank the Ecosystems and Landscape Evolution Group at WSL and ETH Zurich for feedback, B. Flück for technical support, and M. Dawes for editing the manuscript for English. Funding: This work was supported by the Swiss National Science Foundation (grant 310030_188550 to L.P. and supporting A.S. and W.B.); ETH Zürich (internal grant ETH-34 18-1 to L.P.) and postdoctoral fellowship 18-2 FEL-52 to L.M.B.); the German Research Foundation (grant DFG-FZT 118, 202548816 to O.H.); the Swiss National Science Foundation (Postdoc Mobility Fellowship 206844 to I.M.); the European Union’s Horizon 2020 Research and Innovation Programme (Marie Skłodowska-Curie grant 896323 to O.J.R.); and the Prinzessin Therese von Bayern Foundation (E.M.J.). Author contributions: Conceptualization: A.S., L.M.B., I.M., E.M.J., O.H., O.J.R., W.B., V.B., T.K., W.J., L.P.; Funding acquisition: L.P.; Methodology: A.S., L.M.B., I.M., O.H., V.B., T.K.; Investigation: A.S., L.M.B.; Project administration: A.S.; Supervision: L.P.; Visualization: A.S.; Writing – original draft: A.S.; Writing – review & editing: A.S., L.M.B., I.M., E.M.J., O.H., O.J.R., W.B., V.B., T.K., W.J., L.P. Competing interests: The authors declare no competing interests. Data and materials availability: All data generated and analyzed during this study and the code required to replicate the analyses are available in the manuscript or have been deposited at Dryad (49). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf7122 Materials and Methods Figs. S1 to S13 Tables S1 to S3 References (50–143) 33. E. M. Joyce, K. R. Thiele, J. W. F. Slik, D. M. Crayn, Biol. J. Linn. Soc. Lond. 132, 374–387 (2021). Submitted 9 November 2022; accepted 1 June 2023 10.1126/science.adf7122 Skeels et al., Science 381, 86–92 (2023) 7 July 2023 6 of 6
10.1126_science.adg4084
RES EARCH R E S E A R C H A R T I C L E ◥ PHYSICS An improved bound on the electron’s electric dipole moment Tanya S. Roussy1,2†, Luke Caldwell1,2†, Trevor Wright1,2, William B. Cairncross1,2‡, Yuval Shagam1,2§, Kia Boon Ng1,2, Noah Schlossberger1,2, Sun Yool Park1,2, Anzhou Wang1,2, Jun Ye1,2, Eric A. Cornell1,2* The imbalance of matter and antimatter in our Universe provides compelling motivation to search for undiscovered particles that violate charge-parity symmetry. Interactions with vacuum fluctuations of the fields associated with these new particles will induce an electric dipole moment of the electron (eEDM). We present the most precise measurement yet of the eEDM using electrons confined inside molecular ions, subjected to a huge intramolecular electric field, and evolving coherently for up to 3 seconds. Our result is consistent with zero and improves on the previous best upper bound by a factor of ~2.4. Our results provide constraints on broad classes of new physics above 1013 electron volts, beyond the direct reach of the current particle colliders or those likely to be available in the coming decades. E lectric dipole moments of fundamental particles, such as the electron, are sig- natures of time-reversal symmetry viola- tion, equivalent to violation of combined charge and parity (CP) symmetry (1). CP symmetry is broken in the standard model but only in the quark sector (2), so the coupling to leptons is weak and the predicted electron’s electric dipole moment (eEDM) several orders of magnitude below current experimental sen- sitivity (3, 4). Explaining the imbalance of mat- ter and antimatter in the Universe requires additional CP violation, beyond that present in the Standard Model (5–7). Many proposed extensions predict new particles at energies higher than any so far discovered, with CP- violating interactions. These new particles can induce a much larger eEDM, often within reach of near-term experiments (8–10). A non- zero measurement at current experimental sensitivities would unambiguously signal new physics, whereas a more precise measurement consistent with zero imposes challenging constraints on possible explanations of the matter-antimatter imbalance. Our measure- ment uses quantum projection–noise-limited spectroscopy on samples of hundreds of mo- lecular ions with interrogation times of up to → → e (cid:4) E 3 s. Our result, de ¼ (cid:2)1:3 T 2:0stat T 0:6syst Þ(cid:3) ð 10(cid:2)30 e cm, is consistent with zero and gives j < 4:1 (cid:3) 10(cid:2)30 e cm at an upper bound of dej 90% confidence. → An eEDM d e ¼ de^s—with ^s a unit vector along the spin of the electron—subject to an → → electric field, E , has an energy of (cid:2)d . The essence of an eEDM search is to measure the energy shift when ^s is aligned with E com- pared with when it is antialigned. The size of the observable shift scales with the size → , and thus many existing (11–13) and pro- of E posed (14–17) eEDM experiments use elec- trons embedded inside polar molecules, where intramolecular electric fields can be ∼105 times larger than what can be directly ap- plied in the lab. These internal electric fields can be aligned in the lab frame by orienting the molecules with modest external electric fields. Our measurement uses HfFþ molecular ions. In an applied electric field of ∼58 V cm(cid:2)1, the 3D1 v ¼ 0; J ¼ 1 Þ “science” state of the mo- lecule is split into a series of doublets (Fig. 1A). In two of these doublets, highlighted in col- or, the molecule is oriented (18); the upper doublet (orange) has the intramolecular axis parallel to the applied field, whereas the lower doublet (blue) is antiparallel. This intramolec- ular axis defines the direction of an effective ð 1JILA, NIST and University of Colorado, Boulder, CO 80309, USA. 2Department of Physics, University of Colorado, Boulder, CO 80309, USA. †These authors contributed equally to this work. ‡Present address: Atom Computing, Berkeley, CA 94710, USA. §Present address: Schulich Faculty of Chemistry, Technion–Israel Institute of Technology, Haifa 3200003, Israel. *Corresponding author. Email: cornell@jila.colorado.edu Fig. 1. Experiment outline. (A) Level structure of the eEDM-sensitive 3D 1 v ¼ 0; J ¼ 1 Þ state. The horizontal axis indicates mF, the projection of the total angular momentum onto the externally applied electric field. The vertical axis indicates the energy of the states. The direction of the electron spin and effective electric field, Eeff, is indicated for each of the states used in the experiment. (B) Schematic of ion trap, composed of eight radial electrodes and a pair of endcap electrodes. (Inset) Fields applied during experimental sequence: the rotating electric bias field, E →0. rot, and the quadrupole magnetic field, B ð → Roussy et al., Science 381, 46–50 (2023) 7 July 2023 1 of 5 (cid:3) (cid:3) (cid:1) (cid:1) (cid:1) (cid:1) i and ↓l We prepare an incoherent mixture of one of the spin states from each doublet, either or ↓uj ↑uj i and ↑l (Fig. 1A), and p 2 pulse to create a coherent then apply a superposition of the two states in each dou- blet. We allow the superpositions to evolve for a variable amount of time, then apply a p 2 pulse to map the accumulated rela- second tive phase between the states in a doublet onto a population difference between those states. We clean out the population in one of (cid:1) (cid:1) (cid:3) (cid:3) the spin states in each doublet, either ↑u=l or (cid:1) (cid:1) ↓u=l , then count the number of ions in the remaining stretched states by state-selectively photodissociating the molecules and detect- ing the resultant Hf+ ions (25). We use the op- posing orientations of the two doublets in the trap to send the Hf+ ions originating from molecules in each doublet to opposite sides of our imaging microchannel plate (MCP) and phosphor screen assembly (26, 27). We then repeat the procedure with the opposite initial Table 1. Summary of systematic shifts and their uncertainties. Data are as presented in (23). All values are in microhertz. Effect Correction Uncertainty Magnetic ..................................................................................................................................................................................................................... ..................................................................................................................................................................................................................... 0.1 → Nonreversing B 0 Stray B fields + distortion of Erot ..................................................................................................................................................................................................................... Berry’s phase ..................................................................................................................................................................................................................... ..................................................................................................................................................................................................................... Rotation-odd axial secular motion Axial fields at harmonics of Erot ..................................................................................................................................................................................................................... Simultaneous doublet spectroscopy ..................................................................................................................................................................................................................... 3.4 3.4 ..................................................................................................................................................................................................................... Imperfect spatial overlap Imperfect imaging contrast ..................................................................................................................................................................................................................... Other ..................................................................................................................................................................................................................... Rotation-induced mF-level mixing ..................................................................................................................................................................................................................... Total ..................................................................................................................................................................................................................... 0.1 < 0:1 3.2 3.5 1.4 0.4 6.9 RES EARCH | R E S E A R C H A R T I C L E electric field, Eeff ≈ 23 GV cm−1 (19–22), acting on the spin of one of the valence electrons. In the presence of a small magnetic field, the two states in a doublet correspond to the spin of this valence electron being aligned or anti- aligned with Eeff. We prepare a coherent super- position of the two spin states and measure the energy difference using Ramsey spec- troscopy. The eEDM will give a contribution to this energy, T2deEeff, with opposite sign in the two doublets. We perform the measure- ment simultaneously on spatially overlapping clouds of ions prepared in each of the dou- blets. The difference between the measured energies is our science signal. Experimental overview Our experimental apparatus is shown sche- matically in Fig. 1B. An overview of the ex- perimental sequence is given here; more details, including an account of improvements made since our earlier result (12), are presented in (23, 24). The sequence begins with produc- tion and radiofrequency trapping of roughly 20; 000 HfFþ ions. To orient the molecules while maintaining confinement, we rotate → the orienting field, E rot, at angular frequency wrot ¼ 2p (cid:3) 375 kHz and perform our spec- troscopy in this rotating frame. We also apply → 0 , to a quadrupole magnetic field gradient, B create a time-averaged effective bias magnetic field, Brot (24). A B C Fig. 2. Example Ramsey data. (A) Detection of Hfþ ions; ions are assigned to the upper or lower doublet based on their position (upper doublet shown in orange, lower doublet in blue). Counts from a thin central swatch where the assignment is ambiguous (shown in gray) are removed. Images shown are averaged over 60 shots of the experiment. (B) Asymmetries for the upper and lower doublet. (C) Fitted sum and difference asymmetries, AS and AD, used to extract mean and difference frequencies, fm and fd. Middle-time data, where the two doublets are out of phase, were collected in this example dataset for illustrative purposes only. Such data contribute very little to the frequency determination and were not collected during the final precision dataset. Roussy et al., Science 381, 46–50 (2023) 7 July 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E spin state. Example data are shown in Fig. 2A. We count the ions on each side of the screen, in each configuration—typically ∼120 ions from each doublet after a 3-s hold time—and from these two measurements construct the two asymmetries, Au and Al, where Au=l ¼ N u=l In N u=l In (cid:2) N u=l Anti þ N u=l Anti ð1Þ Here, N u=l In and N u=l Anti are the number of ions counted, u/l indicates the upper or lower dou- blet, and the subscripts indicate whether we read out the same state that we prepare (In) or the opposite (Anti). We repeat our measurement at different free-evolution times, generating a pair of Ramsey fringes as shown in Fig. 2B. The frequencies of these two fringes are proportional to the energy splitting in the two doublets. The primary contribution to this energy splitting is the Zeeman splitting, 3gF mBBrot, where gF ¼ (cid:2)0:0031 1ð Þ is the g factor of the science state (28). For our typical experimental parameters, this produces fringe frequencies of ∼100Hz. Other effects, including the eEDM, make small modifications to this frequency. Instability of the intensity of the pulsed lasers used for creation and photodissociation of the ions (24) creates considerable noise in the number of ions measured in each shot of the experiment, typically approximately three times as large as the quantum projection– A B Fig. 3. Systematic shifts in the measurement. (A) Shift in fDB caused by nonreversing magnetic quadrupole field, which can be corrected by using the shift in the fB channel, which is ~460 times as large as the fDBshift. (B) Shift in the fB channel caused by deliberately applied second- harmonic electric field E2h with transverse magnetic field, B. Data show variation in shift as angle q2h between E2h and x axis is varied. E2h is ∼250 times as large as that present in the experiment; B is ∼ 14 mG. (Inset) Lissajous figure traced out by total electric field for greatly exaggerated ratio of E2h=Erot; blue and orange show one-half cycle each. The field points in the (cid:2)x direction for more time than in the þx direction. noise limit on the side of the fringe at 3 s. However, these sources of noise, and many others, are common mode between the two doublets—the exact same laser pulses address both clouds of ions—and so the noise in Au and Al is highly correlated. To take advantage of this, we form the sum and difference asymme- tries AS ¼ Au þ Al and AD ¼ Au (cid:2) Al (Fig. 2C). If we take data when the two doublets are close to being in phase, the noise in AD is drastically reduced (27). The two doublets oscillate at slightly different frequencies, fu and fl, owing to a ∼1=230 fractional difference in their magnetic moments, and so during the eEDM dataset we deliberately take our data at a beat. We take two sets of points: the early-time data, when the two doublets are in phase, and the late-time data ∼230 oscillations later, when they come back into phase again. We can control the time of the second beat by varying the strength of the magnetic bias field, Brot. We fit to AS and AD to extract the fm ¼ mean of the two fringe frequencies, Þ. ð 2 fu (cid:2) fl Þ, and their difference, fd ¼ 1 ð 2 fu þ fl 1 We collect Ramsey fringes in 23 ¼ 8 exper- imental states, corresponding to each possi- ble combination of three binary experimental (cid:5) ¼ T1. ~B is the direction of switches, ~B; ~R; ~I → rot , ~R the the magnetic bias field relative to E → rot, and ~I the direction of rotation direction of E → E rot relative to the imaging MCP at the instant of photodissociation, determining which side (cid:4) of the phosphor screen each of the doublets is imaged onto. A set of Ramsey fringes in each of the 8 switch states forms a block. To minimize the effects of experimental drifts, within a block we interleave data collection for the switch states; the first Ramsey time is recorded for all switch states before moving onto the second Ramsey time for each switch state, and so on. We take the 16 fitted frequencies from each block and form 16 linear combinations to give the components of the measured frequencies, which are even or odd under each of the experi- mental switches. Following (29), we label the components with superscripts that denote the switches under which the quantity is odd. For example, our science signal is f DB , the component of the difference frequency that is odd under ~B but even under ~R and ~I (30). The other channels allow us to diagnose system- atics and monitor experimental performance. Over the course of the dataset, we varied a number of other experimental parameters on timescales slower than a block. These include the state we read out at the end of the Ramsey sequence, denoted ~P ¼ T1 and alternated each block; the order in which the switch states are recorded at each Ramsey time, alternated every other block; three different magnitudes of the magnetic bias field, corresponding to mean fringe frequencies of f 0∼77 , 105, and 151 Hz; and reversal of the waveplates that set the lab-frame handedness of the light used for state preparation and readout. During data collection and analysis, we “blinded” our mea- surement of f DB by adding an unknown offset to this channel. The offset was not removed until our systematics search and analysis (23) were complete. Accuracy evaluation To evaluate the accuracy of our measurement, we searched extensively for systematic shifts before data collection; a summary is given in Table 1. In general, we tuned a variety of ex- perimental parameters over ranges that were large compared with those present during data collection, exaggerating any accompany- ing systematic effects, and observed the re- sponse in our data channels. The only shift we could observe directly in the eEDM channel stems from a nonreversing quadrupole mag- netic field and the difference in magnetic moments between the two doublets, caused primarily by the applied electric field mixing the states of the two doublets with higher rotational levels of the molecule. The f B chan- nel provides a direct measurement of the non- reversing magnetic field and allows us to apply ¼ a correction to our science channel, df DB corr f B dgF , where dgF is half the difference between gF the g factors for the upper and lower doublets (Fig. 3A). Before applying any corrections to the science channel, we suppress this system- atic by actively shimming the currents through Roussy et al., Science 381, 46–50 (2023) 7 July 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Summary of our dataset. Cuts have been applied and the uncer- tainty on each block p ffiffiffiffiffi c2 scaled by to account for overscatter. (A) Histo- gram of data. Error bars show standard deviation of bin counts expected from Poisson distribution. The blue line shows normal distribution. (B) Normal probability plot of fDB, showing that the data are consistent with a normal distribution. The gray line shows expected probability for a normal distribution. (C) Variation of central value under different experimental parameters compared with the overall average value of fDB. Here, N is the average number of trapped HfF+ ions per experimental trial during a block. Other panels in (C) are described in the final paragraph of the Experimental Overview section. ¼ 90 nHz. the coils that apply the magnetic bias field to minimize f B. This shimming was so ef- fective that the mean correction that we ap- plied was well below our statistical sensitivity, df DB corr The shimming and correction procedure leaves us susceptible to other possible effects that cause shifts in f B and f DB with a ratio different from dgF . An important example of gF such a shift is the combination of a transverse magnetic field with an electric field oscillating at 2wrot, which was present in our experiment owing to harmonic distortion in the amplifiers driving the trap electrodes. The harmonic dis- tortion causes the electric field, and thus the magnetic moment of the molecule, to spend more time pointing in one spatial direction than the other, giving a nonzero time-averaged Zeeman interaction with a background mag- netic field. This causes shifts in f B but no cor- responding shifts in f DB, where the effect is canceled by the coincident change in the size of the differential magnetic moment owing to the distortion. Figure 3B shows the shift in f B when we deliberately apply a second- harmonic electric field and vary its angle. We shim out the second harmonic on each elec- trode by feeding forward a second-harmonic signal with the opposite phase, suppressing the amplitude by ∼80 dB. We used magnetic shim coils to null the ambient magnetic field at the trap center to <10 mG. The measured sizes of the residual effects were used to com- pute the maximum size of the systematic during our dataset. A full account of all systematic shifts con- sidered is presented in (23). Measuring the eEDM We collected 1370 blocks over about 2 months, corresponding to ∼620 hours of data and ∼108 ion detection events. Each block results in one value of f DB and thus a single measurement of de . The uncertainty on f DB for each block is calculated with only the standard errors on the asymmetries for that block. We applied cuts to the blinded data on the basis of non-eEDM channels that indicated signal quality. Blocks with late-time contrast <0.2 were cut because of a low signal-to-noise ratio, as were blocks containing fitted fringe frequencies that were >3.5s different from the mean fringe frequen- cy for that switch state. After applying cuts, we were left with 1329 blocks with c2 ¼ 1:07 4ð Þ for f DB. Figure 4, A and B, show the distrib- ution of measured f DB values over the 1329 blocks after relaxing the uncertainty for each ¼ 1:035. The of the blocks by a factor of data are consistent with a normal distribution. Our final statistical uncertainty of 22.8 mHz is obtained with these relaxed uncertainties. Based on the number of ions detected in each shot, this uncertainty is ∼30% above the quantum projection–noise limit. More details on how we determine uncertainties are given in (24). ffiffiffiffiffi c2 p Figure 4C shows how the measured value of f DB depends on experimental parameters varied during the dataset; we find no con- cerning dependencies. We removed our blind on 1 November 2022, and obtained a final value for the eEDM- sensitive frequency channel f DB ¼ (cid:2)14:6 T 22:8stat T 6:9syst mHz ð2Þ Þ Dividing by (cid:2)2Eeff ≃ 1:11 (cid:3) 1031 mHz e−1 cm−1 (21, 31), we obtain a value for the eEDM Þ (cid:3) 10(cid:2)30 e cm ð3Þ de ¼ (cid:2)1:3 T 2:0stat T 0:6syst ð sgn gFð h which is consistent with zero within one stan- dard error. The combined statistical and sys- ¼ 2:1 (cid:3) 10(cid:2)30 e cm, tematic uncertainty, sde improves on our previous work (12) by a factor of ∼37, and on the previous state-of-the-art from the ACME collaboration (13) by a factor of ∼2 . This result and that of the ACME collaboration—two measurements using very different experimental platforms with con- trasting sources of systematic shifts—are con- sistent at slightly above one standard error. Discussion We use our result to obtain an upper bound using a folded Gaussian distribution dej j < 4:1 (cid:3) 10(cid:2)30 e cm 90% confidence Þ ð4Þ ð This limit constrains extensions to the Standard Model that predict new sources of CP-symmetry violation to explain the matter- antimatter asymmetry of the Universe (32). Many extensions, including supersymmetry, the two-Higgs model, and left-right symmetric models, generate an eEDM at the one-loop level (33), with magnitude (9) de ∼ ea0a 2 g2 2p sinf CP m2 e M 2 ð5Þ Here, M is the characteristic mass of new par- ticles with effective coupling strength, g, to the electron; f CP is the phase that describes how strongly the interaction violates CP symmetry; me and e are the mass and charge of the elec- tron respectively; and a is the fine-structure constant. Because deºM (cid:2)2 , and because our limit in Eq. 4 is a factor of ∼2:4 smaller than the limit reported in (13), we are sensitive to new ¼ 1:5 times particles with mass that is as large. ffiffiffiffiffiffiffi 2:4 p To estimate the mass reach of our exper- iment, we need to make assumptions for the size of g2 and sinfCP. For the strong force, quantum electrodynamics, and the weak force, g2 ≈ 1; 1=137; and 10(cid:2)6, respectively. For exten- sions to the Standard Model seeking to explain the matter-antimatter asymmetry, the naive ex- ∼ 1. With this assump- pectation is that sinf CP tion, we can interpret our new limit on de as (cid:8) (cid:7) M ≳ g=a1 40 TeV. For new particles with g2 of order a ∼ 1=137, this bound is an order of magnitude greater than the largest-mass particles that can be directly detected at the Large Hadron Collider (34). 2 Roussy et al., Science 381, 46–50 (2023) 7 July 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E So far, we have assumed that CP violation arises purely from de. Diatomic molecules are also sensitive to pseudoscalar-scalar electron- nucleon coupling, CS (35, 36), and we can interpret our measurement as a linear combi- nation, hf DB (cid:3) sgn gFð Þ ¼ (cid:2)2Eeff de þ 2WSCS , ¼ (cid:2)51 kHz (31) is a molecule-specific where WS h structure constant. Assuming that de is zero, we can instead attribute our measurement to CS, and we find CS ¼ (cid:2)1:4 T 2:2stat T 0:7syst ð Þ (cid:3) 10(cid:2)10 ð6Þ Determining rigorous limits on de and CS requires combining the results of two or more measurements using molecules with different ratios of Eeff to WS. Figure S1 shows a combined fit to the results of this work and (13), giving j < 2:1 (cid:3) 10(cid:2)29 e cm and upper bounds of dej j < 1:9 (cid:3) 10(cid:2)9 with 90% confidence. Our CSj measurement improves these bounds by fac- tors of 16 and 12, respectively (37). RE FE RENCES AND N OT ES 1. I. B. Khriplovich, S. K. Lamoreaux, CP Violation Without Strangeness: Electric Dipole Moments of Particles, Atoms, and Molecules, Texts and Monographs in Physics (Springer, 1997). 2. C. Patrignani, Chin. Phys. C 40, 100001 (2016). 3. Y. Yamaguchi, N. Yamanaka, Phys. Rev. Lett. 125, 241802 (2020). 4. Y. Ema, T. Gao, M. Pospelov, Phys. Rev. Lett. 129, 231801 (2022). 5. M. Dine, A. Kusenko, Rev. Mod. Phys. 76, 1–30 (2003). 6. A. D. Sakharov, Pis’ma Z. Eksp. Teor. Fiz. 5, 32–35 7. M. B. Gavela, P. Hernández, J. Orloff, O. Pène, Mod. Phys. Lett. 33. T. Chupp, P. Fierlinger, M. Ramsey-Musolf, J. Singh, A 09, 795–809 (1994). 8. M. Pospelov, A. Ritz, Ann. Phys. 318, 119–169 (2005). 9. J. Engel, M. J. Ramsey-Musolf, U. Van Kolck, Prog. Part. Nucl. Phys. 71, 21–74 (2013). 10. Y. Nakai, M. Reece, J. High Energy Phys. 2017, 31 (2017). 11. J. J. Hudson et al., Nature 473, 493–496 (2011). 12. W. B. Cairncross et al., Phys. Rev. Lett. 119, 153001 (2017). 13. ACME Collaboration, Nature 562, 355–360 (2018). 14. I. Kozyryev, N. R. Hutzler, Phys. Rev. Lett. 119, 133002 (2017). 15. A. Vutha, M. Horbatsch, E. Hessels, Phys. Rev. A 98, 032513 (2018). 16. The NL-eEDM collaboration et al., Eur. Phys. J. D 72, 197 (2018). 17. N. Fitch, J. Lim, E. Hinds, B. Sauer, M. Tarbutt, Quantum Sci. Technol. 6, 014006 (2020). 18. A. E. Leanhardt et al., J. Mol. Spectrosc. 270, 1–25 (2011). 19. E. R. Meyer, J. L. Bohn, M. P. Deskevich, Phys. Rev. A 73, 062108 (2006). 20. A. Petrov, N. Mosyagin, T. Isaev, A. Titov, Phys. Rev. A 76, 030501 (2007). 21. L. V. Skripnikov, J. Chem. Phys. 147, 021101 (2017). 22. T. Fleig, Phys. Rev. A 96, 040502 (2017). 23. L. Caldwell et al., Systematic and statistical uncertainty evaluation of the HfF+ electron electric dipole moment experiment. arXiv:2212.11837 [physics.atom-ph] (2022). 24. Materials and methods are available as supplementary materials. 25. K. K. Ni et al., J. Mol. Spectrosc. 300, 12–15 (2014). 26. Y. Shagam et al., J. Mol. Spectrosc. 368, 111257 (2020). 27. Y. Zhou et al., Phys. Rev. Lett. 124, 053201 (2020). 28. H. Loh, “Search for an electron electric dipole moment with trapped molecular ions,” thesis, University of Colorado Boulder and JILA (2013). 29. J. Baron et al., Science 343, 269–272 (2014). 30. We define the mean frequency to be always positive. If we had instead allowed the fringe frequency to change sign when the sign of the magnetic bias field changes, our science signal would have been fD. 31. T. Fleig, M. Jung, J. High Energy Phys. 2018, 12 (2018). 32. L. Canetti, M. Drewes, M. Shaposhnikov, New J. Phys. 14, Rev. Mod. Phys. 91, 015001 (2019). 34. T. Bose et al. in Proceedings of the 2022 Snowmass Summer Study, Seattle, WA, 17 to 26 July 2022. 35. T. Chupp, M. Ramsey-Musolf, Phys. Rev. C Nucl. Phys. 91, 035502 (2015). 36. M. Denis et al., New J. Phys. 17, 043005 (2015). 37. Improvements are given with respect to equivalent bounds calculated from combined fit to results of (13) and (12). 38. T. S. Roussy et al., Data presented in “An improved bound on the electron’s electric dipole moment,” Zenodo (2023); https://doi.org/10.5281/zenodo.7837398. AC KNOWLED GME NTS We thank the staff at JILA for making this experiment possible. Funding: This work was supported by the Cydney and Tom Marsico Family Foundation, the Sloan Foundation, the Gordon and Betty Moore Foundation, NIST, and the NSF (award PHY-1125844). T.W. acknowledges funding support from NSF GRFP. Author contributions: All authors contributed, in varying degrees, to the conception of experimental approach, molecular survey spectroscopy, apparatus design and construction, apparatus debugging and commissioning, apparatus maintenance, data collection and analysis, analysis of systematic errors, and preparation of publications. We are all responsible for the accuracy of the bottom-line result and of the written account we have submitted to Science. Competing interests: The authors declare that they have no competing interests. Data and materials availability: Data and code used for analysis are available at Zenodo (38). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg4084 Materials and Methods Figs. S1 and S2 References (39–42) Submitted 22 December 2022; accepted 18 May 2023 10.1126/science.adg4084 (1967). 095012 (2012). Roussy et al., Science 381, 46–50 (2023) 7 July 2023 5 of 5
10.1126_science.adg7693
RES EARCH R E S E A R C H A R T I C L E ◥ FRACTURE MECHANICS Tensile cracks can shatter classical speed limits Meng Wang, Songlin Shi, Jay Fineberg* Brittle materials fail by means of rapid cracks. Classical fracture mechanics describes the motion of tensile cracks that dissipate released elastic energy within a point-like zone at their tips. Within this framework, a “classical” tensile crack cannot exceed the Rayleigh wave speed, cR. Using brittle neo- hookean materials, we experimentally demonstrate the existence of “supershear” tensile cracks that exceed shear wave speeds, cR. Supershear cracks smoothly accelerate beyond cR, to speeds that could approach dilatation wave speeds. Supershear dynamics are governed by different principles than those guiding “classical” cracks; this fracture mode is excited at critical (material dependent) applied strains. This nonclassical mode of tensile fracture represents a fundamental shift in our understanding of the fracture process. H ow do materials break? How fast can a crack propagate? These related and fun- damental questions are of essential in- terest to a broad range of scientific and engineering communities. In tensile fracture, the maximal velocity of a moving crack is classically considered to be (1) the Rayleigh wave speed, cR , although there have been suggestions that cracks could surpass this speed (2, 3). In the vicinity of a crack’s tip, remotely applied stresses are amplified to an approximate singularity. Crack motion is guided by the principle of energy balance. This principle states that fracture takes place when the flux of stored potential energy flow- ing from large (system-size) scales to the crack’s tip balances the material’s fracture energy, G.G, the energy dissipated at the tip, is defined as the energy per unit crack advance required to separate the material. Provided that dissipa- tion is confined to this point-like zone and crack instabilities are suppressed, fracture mechanics provide the fundamental theoret- ical framework to describe crack dynamics (1, 4, 5). Crack velocities, v, smoothly accelerate to cR while maintaining energy balance (6). Beyond cR, fracture mechanics predict that the energy flux into a crack’s tip will become negative, rendering v > cR to be unphysical. An exception to this speed limit could oc- cur in the case of cracks driven by shear load- ing (mode II) (7, 8). Analytical solutions of mode II cracks having a finite-sized dissipa- tive zone (9) predict positive energy flux into “supershear” cracks moving with speeds be- tween the shear wave speed, cs , and the di- latational wave speed, cp . Such rapid shear cracks are observed in interfacial and fric- tional failure (10, 11), as well as in earthquakes (12, 13). When sufficient elastic energy exists, shear cracks approaching cR may transition to supershear by giving birth to a daughter crack that moves faster than cs (14–16), thereby Racah Institute of Physics, Hebrew University of Jerusalem, Jerusalem 91904, Israel. *Corresponding author. Email: jay@mail.huji.ac.il ð Fig. 1. Experimental setup and examples of subsonic and supershear cracks. (A) Cracks propagating along unweakened (left) and weakened (middle) materials driven through applied uniaxial stretch, l ¼ h þ Dh Þ=h (Dh is the imposed constant displacement in y). Deformation fields are measured via distortions of 80-mm grids, imprinted on one xy surface at z ¼ 0. (Right) Grooves along y ¼ 0 form straight weak layers by reducing material thicknesses from W0 to W. (B) Photographs of cracks propagating at v ¼ 0:68cR along unweakened (left, l ¼ 1:10) W ¼ W0 and weakened (right, l ¼ 1:07) W=W0 ¼ 0:5 materials demonstrate parabolic CTODs at the millimeter scale (dashed lines). Black regions along weak layers are caustics (diverging light paths) caused by the highly curved boundaries of weak layers. (C) Measured energy flux, G vð Þ, for W=W0 ¼ 0:5 (red) and W=W0 ¼ 1 [blue; data taken from (6, 28)]. G vð Þ was measured via either the CTOD (open circles) or the J-integral calculated with measured deformation fields (filled circles). G vð Þ scales linearly with W=W0. (D) Supershear cracks in samples without (left, l ¼ 1:36) and with (middle, l ¼ 1:34) a weak layer. (Right) Magnified views of areas denoted by dashed rectangles highlight discontinuous deformation fronts formed by shock waves emanating from crack tips in supershear (v > cs) propagation. Wang et al., Science 381, 415–419 (2023) 28 July 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E extending the dissipative region from a point to a finite region in space. The motion of such supershear cracks can be approximately de- scribed through the principle of energy bal- ance (17). Cracks driven by the tensile loading (mode I) are commonly believed to be limited by cR (18). This limit, as well as the corresponding equa- tion of motion predicted by linear elastic frac- ture mechanics (LEFM), has been confirmed experimentally (6, 19, 20). Extensions of clas- sical fracture to hyperelastic materials predict that tensile cracks could surpass cR (21). A qual- itatively different mode of supershear tensile fracture not incorporated in classical fracture mechanics has been predicted by lattice models to occur at high applied stretch (22–24). This theory predicts cracks that are able to exceed cs and even cp, regardless of material stiffening or softening (25). Moreover, these states are not expected to be governed by the principle of energy balance, the cornerstone of the classi- cal theory of fracture. In rubber, experiments have observed marginal supershear propaga- tion under extreme stretches (2, 3); however, both unequivocal experimental evidence and fundamental understanding of supershear ten- sile fracture are still lacking. Experimental system Our experiments were performed in thin, strip- like sheets of polyacrylamide hydrogels having varied heights, h, where x; y; z are the propa- gation, loading, and sheet thickness directions (Fig. 1A). Polyacrylamide gels are brittle neo- hookean materials: linearly elastic at small stretches, l≈1, and nonlinearly elastic as l is increased. LEFM describes the motion of straight single cracks in these materials (6) for crack speeds v < cR. The elastic nonlinearity Fig. 2. Typical experiments with and without a weak layer. (A) Crack speed v, as a function of crack lengths l in x, for cracks propagating along weak layers under different applied stretches. Sample height h = 20 mm and W=W0 ¼ 0.5. The values of cR and cs are shown by the gray dashed and solid line, respectively. Colored dashed lines denote va. (B) Photographs of supershear cracks (top) and the strain energy density (SED) fields (in J/m3) around the crack tip in the reference frame (bottom) at different values of va corresponding to the four supershear cases shown in (A). (C) Mach cones angles, a, measured from the SED fields in the reference frame, as a function of va=cs, with cs measured from the shear modulus. a of the four cases shown in (A) and (B) are denoted by the colors in (A). (D) Examples of dynamics of supershear oscillatory cracks. Orange markers show, as an example, the crack speed component in x, vx for l = 1.26, attaining an asymptotic value highlighted by the orange dashed line. (E) Three snapshots of supershear oscillatory cracks, whose speeds are noted by the full markers in (D), at l = 1.50. in the near-tip region (26, 27) causes oscilla- tory crack motion when v→~0.9cs. In Fig. 1A, we experimentally show that crack oscillations and microbranching instabilities (4) can be suppressed by guiding cracks through a straight weak layer (at y ¼ 0 ) that constrains crack motion to straight paths (materials and meth- ods). The deformation fields of the material surrounding crack tips, for both straight (weak layer) and oscillatory (no weak layer) cracks, were measured by imprinting an initially square grid mesh on one (z ¼ 0) of the sample’s xy faces. During crack propagation, we recorded the grid’s instantaneous deformation with a high-resolution fast camera (Fig. 1B). For subsonic crack propagation at small stretches, G vð Þ, the instantaneous energy flux from the effectively two-dimensional medium into the crack tip (the “energy release rate”) was measured by using both the grid defor- mation and the crack tip’s opening displace- ment (6, 28). When a weak layer is used, the strain in the loading direction within the weak layer is larger than within the outside region by a factor of W0=W, where W =W0 is the thickness ratio of the fractured region (fig. S1 and supplementary text section 1). The strain energy driving fracture is, however, mainly supplied by the outer region. As a result, in samples with weak layers, G vð Þ is propor- tional to W =W0 (Fig. 1C); the effective energy dissipation of cracks within weak layers is ~G vð Þ ¼ G vð ÞW =W0, where G vð Þ is the fracture energy of the material. No additional effects of the weak layer on crack dynamics or struc- ture were observed. For low l, straight crack dynamics are well-described by LEFM (6), ac- celerating smoothly until cR; increasing l sim- ply increases acceleration toward cR. Observation of supershear cracks for large stretches At higher l (beyond ∼1.2), crack dynamics change substantially; supershear cracks appear. Figure 1D presents typical examples of both supershear oscillatory cracks and straight super- shear cracks that are guided by weak layers (movies S1 to S3). Deformation fields sur- rounding supershear cracks radically differ from sub-Rayleigh (v < cR) cracks. Supershear cracks have wedge-like opening displacements (23, 29) and shock waves emanating from their tips. At a shock wave, deformation fields discontinuously jump. Behind these shocks, deformation fields are highly distorted, and the kinetic energy density (fig. S2) increases precipitously relative to the strain energy den- sity with v. Ahead of the shock, variations of the deformations, as well as the kinetic and strain energy densities, are much smaller. Supershear crack speeds, v, versus crack lengths,l, for variousl are presented in Fig. 2A, for propagation along a weak layer. Cracks initiate at subsonic speeds and accelerate Wang et al., Science 381, 415–419 (2023) 28 July 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Strain field dependence of supershear cracks. (A and B) Measurement of the Dxx strain component of supershear cracks propagating at (A) va ¼ 1:04cs (l ¼ 1:22) and (B) va ¼ 1:28cs (l ¼ 1:56). Dashed lines denote angular locations of the strain measurements. r and q represent the distance to the crack tip and the angle relative to x, respectively. (C) Radial dependences of the strain fields for va ¼ 1:04cs ahead, Dp xx (right, inset) the Mach cone. xx (left, inset) and behind Ds ð xx (left) and Ds xx (right) when normalized by the far-field values Measurements were performed at the values of q denoted by the colors in (A). Main panels: Dp Dxx r ¼ 5mm; q (D) Radial dependences of Dp xx qð Þ (left, inset) and Ds xx Main panels demonstrate the collapse of these functions, when normalized as in (C). Colors are the same as for the dashed lines in (B). Þ for each q. The black dashed line (right) depicts a 1=r dependence. qð Þ (right, inset) at va ¼ 1:28cs. smoothly until reaching l-dependent asymp- totic speeds, va (movie S1). This continuous transition is very different from the transition to supershear of shear cracks, which both nucleate at finite distances ahead of sub- Rayleigh cracks, and the range cR < v < cs is forbidden (7–9). Figure 2B shows the grid de- formations and corresponding strain energy densities surrounding crack tips for 1:04cs ≤ v ≤ 1:31cs. In the reference (material) frame, the shock waves generated form Mach cones angles, a, relative to x, where sin að Þ ¼ cs=va (Fig. 2C). Shear wave speeds derived from the Mach cones, cs = 5.75 T 0.22 m/s, agree well with both values calculated from the shear modulus,m, and directly measured values (mate- rials and methods). The speed of waves prop- agating in a stretched medium was determined to be accurately described by the neo-hookean constitutive law of the material in our experi- ments, with a constant value of cs (fig. S3 and supplementary text section 2). Without weak layers, supershear oscillatory cracks appear at the same l levels (Fig. 2D) as the straight supershear examples presented in Fig. 2A. Here, whereas the global speed, v, oscillates with increasing amplitudes, the velocity com- ponent in x, vx, reaches a constant (asymptotic) speed, va = vx, before the oscillation develops (fig. S4 and supplementary text section 3). Oscillatory supershear cracks generate shock waves that, in contrast to Mach cones of straight cracks, have irregular shapes that vary in time (Fig. 2E) as they reflect the spatially oscillating crack tips. ð xx r; qð For supershear cracks, Dxx is the strain component with the largest variations near the crack tip. In Fig. 3, A and B, we present measurements of Dxx for va = 1.04cs and 1.28cs. We first consider the (P-wave governed) fields ahead of the Mach cones, which we denote as Dp xx. For both velocities, Fig. 3, C and D (left), show that Dp Þ have very different radial dependences for each q. When, however, nor- Þ col- malized by Dp lapse onto well-defined radial (v-dependent) Þ. This collapse suggests that functions, fr r; vð Þ xx r; qð xx are separable functions of the form Dp Dp ≈fr r; vð Þ. This separable nature is also true for Dp xy . Although neither of the functions fr r; vð Þ possesses a pure power-law Þ in- form, near the transition, fr r; v ¼ 1:04cs creases sharply as r approaches the shock front, whereas at the larger velocity, fr r; v ¼ 1:28csÞ is significantly “less” singular. xx r ¼ 5mm; q yy and Dp Þ (cid:2) fq q; vð Þ, all Dp xx r; qð We now consider the (shear wave–dominated) strain fieldsDs xx behind the Mach cone (Fig. 3, C and D, right panels). The same normalization for ð ð va = 1.04cs produces an approximate collapse to a ∼1=r form for large r, but does not collapse the strain fields for r < 2 mm. At the higher Þ do approximately collapse to velocity, Ds the form Ds Þ (Fig. 3D, Þ (cid:2) gq q; vð Þ≈gr r; vð right), with a nearly linear nonsingular decay of gr that is much weaker than the decay of Ds xx for va = 1.04cs. This qualitative behavior is echoed by the Ds xx r; qð xx r; qð xy components. yy and Ds The transition to supershear cracks and their dynamics are governed by l The energy released by supershear cracks, Ga, is measured at va in the strip geometry. At vx ¼ va, translational invariance in x exists, so for both straight and oscillatory supershear cracks, the strain energy density, we lð Þ (we is defined in supplementary text section 5), is constant far ahead of the crack tip. Hence, the energy released per unit crack advance is Ga ¼ we lð Þ (cid:2) h. For v < cR, Ga ¼ G vð Þ ¼ ~G vð Þ (1). For v < cR, the values of G vð Þ measured via the J-integral, crack tip opening displace- ment (CTOD), and strip methods are identical (supplementary text section 5). In Fig. 4A, we triggered supershear cracks for various values of Ga by varying h and l. Although va varies with Ga for a given h, va is not a universal function of Ga; variation of h Wang et al., Science 381, 415–419 (2023) 28 July 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E produces different va for the same values of Ga. Thus, in contrast to sub-Rayleigh fracture, Fig. 4A demonstrates that the “classical” en- ergy balance [Ga ¼ ~G vð Þ] is not the governing principle for supershear dynamics. What drives cracks to be supershear? Figure 4B indicates that the supershear transition takes place at a critical value of l ¼ lt ¼ 1:19 T 0:01, for the gel composition used. In particular, the dependence of va on l remains Fig. 4. Dynamics of supershear cracks. (A) Asymptotic crack speeds va as a function of Ga for three different system heights, h, of 20, 40, and 80 mm with W=W0 ¼ 0:5 while varying l. Each data point corresponds to a single experiment. cR and cs are denoted by the gray dashed and solid line, respectively. When h is varied, each va corresponds to different values of Ga. This breakdown of the collapse of crack dynamics with G (that is an inherent property of LEFM) already takes place within the transition region (cR < v < cs) to supershear. (B) Asymptotic crack speeds va as a function of l. Symbols with the colors in (A) represent the same experiments. Purple and black symbols correspond to straight supershear cracks with different values of W=W0 for h ¼ 20 mm. All va collapse to a unique function of l (including va within the transition region), regardless the value of W=W0. Supershear oscillatory cracks (red) collapse onto the same curve. Representative error bars (SD) for oscillatory cracks are shown. Orange dashed line denotes the value of l ¼ 1:19 T 0:01 at the supershear transition. Fig. 5. Dependence of supershear cracks on the chemical composition of gels. (A) Asymp- totic crack speeds, va, as a function of l for gels with different concentrations. Gray dashed and solid lines correspond to cR and cs, respectively. Orange symbols are the data shown in Fig. 4B. Supershear cracks appear at increased values of lt when either the crosslinker concentration is decreased (green and purple) or the amount of monomer is increased (blue). When crosslinker and monomer concentrations are proportionally varied, va versus l is invariant (red, orange). (B) Strain at the supershear transition, lt (cid:3) 1, for gels of different monomer-to- crosslinker molar ratios, M. (lt values were determined by a spline interpolation at va=cs = 1). Symbol colors correspond to the respective gels in (A). A representative error bar (SD) is presented. The black dashed line, with a slope of 0.5, is a guide to the eye. (Inset) A linear fit of lt (cid:3) 1 with (red dashed line) produces a slope of 0.021 with no intercept. (C) Collapse of va versus l curves for gels of different chemical compositions when l is normalized by lt. A transition in the va versus l relation appears at va ecR. ffiffiffiffi M p invariant even when ~G vð Þ is changed by vary- ing W =W0 . Even without a weak layer, the same critical value of lt ¼ 1:19T0:01 governs the transition from subshear to supershear oscillatory cracks. Moreover, Fig. 4B demon- strates that va depends solely on l; for a given gel, all supershear velocities, both for straight and oscillatory cracks, collapse to a single, well-defined function of l. No effects of the geometry and weak layer height on the dynam- ics of supershear cracks were observed (fig. S5 and supplementary text section 4). Supershear dynamics are governed by the stretch level, l, not by Ga. G must still be greater than G csð Þ for super- shear fracture to take place. Even if l > lt, sub- shear propagation will occur when h is so small that G 0ð Þ < Ga < G csð Þ (supplementary text section 5 and figs. S6 and S7). As Ga > G csð Þ is necessary, a critical strip height hc ¼ G csð Þ=we ltð Þ, is required to support super- shear. For the gels used in Fig. 4, hc≈ 12.8 mm for W ¼ W0. hc decreases proportionally with W =W0. What controls the value of lt? To address this question, we performed experiments on samples composed of varying monomer and crosslinker concentrations (Fig. 5). M, the monomer-to-crosslinker molar ratio, deter- mines the number of polymer segments be- tween crosslinks (supplementary text section 6). Figure 5A shows that both lt and the overall dependence of va on l are unaffected when monomer and crosslinker concentra- tions are increased proportionally (thus fix- ing M while appreciably varying both the elastic moduli and G; table S1). When, how- ever, M is changed by separately varying the monomer or crosslinker concentrations, the relation between va and l systematically var- ies. In particular, Fig. 5B shows that the strain at the transition, lt (cid:3) 1, scales approximately with a proportionality factor, linearly with a∼0:021. Figure 5C demonstrates that nor- malizing l by lt collapses all of the va − l relations for different gels to a single curve with a sharp transition from subsonic cracks to the supershear branch at l=lt ¼ 1. The slight divergence for gels with high mono- mer concentrations may be due to extensive polymer entanglement, which also produces high G (30). ffiffiffiffiffi M p Discussion Figure 5B demonstrates that lt , the macro- scopic critical stretch at the supershear tran- sition, depends critically on a microscopic quantity, M, that characterizes the gels’ in- ternal structure. Fracture mechanics enables us to relate the applied stretch to a micro- scopic scale, the cohesive zone size, dt . We will now show that this connection provides ffiffiffiffiffi us with a way to understand the lt (cid:3) 1º M scaling presented in Fig. 5B. p Wang et al., Science 381, 415–419 (2023) 28 July 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E We consider the simplest cohesive zone model, the Dugdale model (1) dt ¼ G csð Þ sc ¼ Þhc we ltð sc ð1aÞ where sc is the maximal stress that the mate- rial can sustain. Replacing we ltð Þ in Eq. 1a by its leading order term, 3m=2ðlt (cid:3) 1Þ2, and sub- stituting the scaling result suggested by Fig. 5B, lt (cid:3) 1 ≈ a , yields ffiffiffiffiffi M p p ffiffiffiffiffi M ð1bÞ dt ≈ 3ma2hcM 2sc Equation 1b predicts that dt is proportional to d ¼ Mb=2, the unfolded (stretched) polymer length between two crosslinks. The propor- tionality factor is 3ma2hc= scbð Þ, where b is the length of a monomer unit. The fact that dt and d scale in the same way with M and 3ma2hc= scbð Þ ∼ O 1ð Þ (table S1 and supplementary ma- terial section 6) suggests that these two os- tensibly independent quantities are related. p ffiffiffiffiffi M Hence, the scaling of lt (cid:3) 1 with sug- gests a transition of the polymer structure within the cohesive zone, from coiled polymer chains (whose lengths scale as ) to stretched chains whose strength, sc , is determined by the internal forces within the polymer chains (31). This transition may both explain the heu- ristic scaling presented in Fig. 5B and could provide a clue toward a physical picture of the transition from classical fracture to supershear. Fracture mechanical solutions for supershear tensile cracks do exist in the literature (1), al- though they are considered to be nonphysical on energy grounds. These solutions exhibit qualitative behavior that is very different from what we experimentally observe here. Both ahead and behind the Mach cone, the theo- retical solutions are singular (∼1=rq), with the same singularity, q vð Þ. Moreover, these solu- tions predict q vð Þ to be q ∼ 0 at the transition to supershear with an increase to a maximal value of q ¼ 1=2 at v ¼ cs. By contrast, our experiments clearly show that the observed supershear cracks have very different radial dependences on both sides of their Mach cones. In addition, although the measured fields are not truly singular, they have their strongest r dependence when propagating at speeds close to cs. Beyond cs, the fields’ radial dependence then weakens with increasing v. There also exist numerical solutions for supershear cracks in hyperelastic materials in which the local wave speed within the highly stretched region at the crack tip is higher than the wave speeds in the surrounding bulk mate- rial (21). In such cases, cracks are “locally” sub- sonic, whereas crack velocities appear to be beyond cs to material far from the crack tip. Such effects have been recently observed in shear fracture (32). The supershear branch described in our experiments does not ap- ffiffiffi 2 p pear to correspond to these types of cracks because, at the strains applied in our experi- ments, the gels used are well-described by neo- hookean constitutive relations—even within the weak layers up to l≈ 1.8 (fig. S1). In neo- hookean materials, it is known that cs is in- variant with stretch (23, 28), and we have verified this with direct measurements (sup- plementary text section 2 and fig. S3). Our results therefore suggest that, as Marder (23) predicted, an entirely different branch of solutions for propagating cracks exists. More- over, these results suggest that the supershear solution branch is both entirely general and independent of the microscopic material struc- ture; the elastic gels in the experiments and the brittle lattices in the numerics possess wholly different microscopic properties. Even if the experimentally observed supershear states cor- respond to the branch of solutions observed in these lattice models (22–25), many important fundamental questions arise that challenge our fundamental understanding of fracture. For example, our results imply that the tran- sition mechanism to supershear takes place within the small scales surrounding the crack tip. The polymer-stretching transition sug- gested here is, however, certainly not realized in lattice models. As such, it is unclear what general mechanisms within the dissipative re- gion surrounding the crack tip give rise to the supershear transition. There are also puzzles at large (system-size) scales. In contrast to supersonic cracks driven by extreme (explosive) loading rates (33), the supershear cracks described here are driven by the release of elastic energy that is stored within the macroscopic scales that are roughly defined by hc. hc is both insensitive to local stretch along the crack path and is a scale that is much larger than both the crack tip region and/or the weak layer scale. An important question is, therefore, by what mechanism does the material supporting these rapidly propagat- ing states self-organize, to transport the strain energy within h≤hc to the dissipative region at supershear crack tips? This mechanism must be wholly different from that described by clas- sical fracture mechanics; a “classical” crack is unable to transport energy to its tip for v≥cR. RE FERENCES AND NOTES 1. L. B. Freund, Dynamic Fracture Mechanics (Cambridge Univ. Press, 1998). 2. P. J. Petersan, R. D. Deegan, M. Marder, H. L. Swinney, Phys. Rev. Lett. 93, 015504 (2004). 3. T.-T. Mai, K. Okuno, K. Tsunoda, K. Urayama, ACS Macro Lett. 9, 762–768 (2020). 4. E. Sharon, J. Fineberg, Nature 397, 333–335 5. (1999). I. Svetlizky, D. S. Kammer, E. Bayart, G. Cohen, J. Fineberg, Phys. Rev. Lett. 118, 125501 (2017). 6. T. Goldman, A. Livne, J. Fineberg, Phys. Rev. Lett. 104, 114301 (2010). 7. D. Andrews, J. Geophys. Res. 81, 5679–5687 (1976). 8. R. Burridge, G. Conn, L. Freund, J. Geophys. Res. 84, 2210 (1979). 9. K. B. Broberg, Cracks and Fracture (Elsevier, 1999). 10. A. J. Rosakis, O. Samudrala, D. Coker, Science 284, 1337–1340 (1999). 11. F. X. Passelègue, A. Schubnel, S. Nielsen, H. S. Bhat, R. Madariaga, Science 340, 1208–1211 (2013). 12. M. Bouchon, M. Vallée, Science 301, 824–826 (2003). 13. E. M. Dunham, R. J. Archuleta, Bull. Seismol. Soc. Am. 94, S256–S268 (2004). 14. A. Needleman, J. Appl. Mech. 66, 847–857 (1999). 15. F. F. Abraham, H. Gao, Phys. Rev. Lett. 84, 3113–3116 (2000). 16. K. Xia, A. J. Rosakis, H. Kanamori, Science 303, 1859–1861 (2004). 17. D. S. Kammer, I. Svetlizky, G. Cohen, J. Fineberg, Sci. Adv. 4, eaat5622 (2018). 18. K. Broberg, in Structural Integrity (Springer, 1989), pp. 1–13. 19. P. D. Washabaugh, W. Knauss, Int. J. Fract. 65, 97–114 (1994). 20. T. Cramer, A. Wanner, P. Gumbsch, Phys. Rev. Lett. 85, 788–791 (2000). 21. M. J. Buehler, F. F. Abraham, H. Gao, Nature 426, 141–146 (2003). 22. L. I. Slepyan, Sov. Phys. Dokl. 260, 900 (1981). 23. M. Marder, J. Mech. Phys. Solids 54, 491–532 (2006). 24. Y. Jia, W. Zhu, T. Li, B. Liu, J. Mech. Phys. Solids 60, 1447–1461 (2012). 25. T. M. Guozden, E. A. Jagla, M. Marder, Int. J. Fract. 162, 107–125 (2010). 26. C.-H. Chen, E. Bouchbinder, A. Karma, Nat. Phys. 13, 1186–1190 (2017). 27. Y. Lubomirsky, C.-H. Chen, A. Karma, E. Bouchbinder, Phys. Rev. Lett. 121, 134301 (2018). 28. A. Livne, E. Bouchbinder, I. Svetlizky, J. Fineberg, Science 327, 1359–1363 (2010). 29. S. Moulinet, M. Adda-Bedia, Phys. Rev. Lett. 115, 184301 (2015). 30. J. Kim, G. Zhang, M. Shi, Z. Suo, Science 374, 212–216 (2021). 31. M. Rubinstein, R. H. Colby et al., Polymer Physics (Oxford Univ. Press, 2003). 32. M. Gori, V. Rubino, A. J. Rosakis, N. Lapusta, Nat. Commun. 9, 4754 (2018). 33. S. Winkler, D. Shockey, D. Curran, Int. J. Fract. Mech. 6, 151–158 (1970). 34. M. Wang, (2023). Data for: Tensile cracks can shatter classical speed limits, Dryad. AC KNOWLED GME NTS We thank M. Adda-Bedia (ENS Lyon) and D. Kammer (ETH Zurich) for helpful discussions. Funding: This work was supported by Israel Science Foundation (grant no. 840/19). M.W. acknowledge the support of the Lady Davis Fellowship Trust. Author contributions: M.W. and J.F. conceived of the project. M.W. designed and performed research. J.F. assisted with experimental design. M.W., S.S., and J.F. analyzed the results. M.W. and J.F. wrote the manuscript. J.F. supervised the research. All authors discussed the result and commented on the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: All data are available in the manuscript or the supplementary materials and are available at the Dryad repository (34). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg7693 Materials and Methods Supplementary Text Figs. S1 to S7 Table S1 References (35–41) Movies S1 to S3 Submitted 19 January 2023; accepted 2 June 2023 10.1126/science.adg7693 Wang et al., Science 381, 415–419 (2023) 28 July 2023 5 of 5
10.1126_science.adg8155
RES EARCH R E S E A R C H A R T I C L E ◥ PHYSICS Quantum control of trapped polyatomic molecules for eEDM searches Loïc Anderegg1,2*, Nathaniel B. Vilas1,2, Christian Hallas1,2, Paige Robichaud1,2, Arian Jadbabaie3, John M. Doyle1,2, Nicholas R. Hutzler3* Ultracold polyatomic molecules are promising candidates for experiments in quantum science and precision searches for physics beyond the Standard Model. A key requirement is the ability to achieve full quantum control over the internal structure of the molecules. In this work, we established coherent control of individual quantum states in calcium monohydroxide (CaOH) and demonstrated a method for searching for the electron electric dipole moment (eEDM). Optically trapped, ultracold CaOH molecules were prepared in a single quantum state, polarized in an electric field, and coherently transferred into an eEDM-sensitive state where an electron spin precession measurement was performed. To extend the coherence time, we used eEDM- sensitive states with tunable, near-zero magnetic field sensitivity. Our results establish a path for eEDM searches with trapped polyatomic molecules. T he rich structure of polyatomic molecules makes them an appealing platform for experiments in quantum science (1–4), ultracold chemistry (5), and precision mea- surements (6–10). Key to this structure is the presence of near-degenerate states of oppo- site parity, which allow the molecules to be easily polarized in the laboratory frame with the application of a small electric field. Such states are generic among polyatomic mole- cules, but rare in diatomics, and may be useful for applications such as analog simulation of quantum magnetism models (1, 2) or for real- izing switchable interactions and long-lived qubit states for quantum computing (4). Addi- tionally, the parity-doublet states in trapped polyatomic molecules are expected to be an invaluable tool for systematic error rejection in precision measurements of physics beyond the Standard Model (BSM) (6). To date, several species of polyatomic molecules have been laser cooled and/or trapped at ultracold tem- peratures (11–17). One powerful avenue for tabletop BSM searches is probing for the electron electric dipole moment (eEDM) (18–22), de, which violates time-reversal (T) symmetry and is predicted by many BSM theories to be orders- of-magnitude larger than the Standard Model prediction (19, 20). Present state-of-the-art eEDM experiments are broadly sensitive to T-violating physics at energies much greater than 1 TeV (23–28). All such experiments use Ramsey spectroscopy to measure an energy shift caused by the interaction of the electron 1Department of Physics, Harvard University, Cambridge, MA 02138, USA. 2Harvard-MIT Center for Ultracold Atoms, Cambridge, MA 02138, USA. 3Division of Physics, Mathematics, and Astronomy, California Institute of Technology, Pasadena, CA 91125, USA. *Corresponding author. Email: anderegg@g.harvard.edu (L.A.); hutzler@caltech.edu (N.R.H.) with the large electric field that is present in- side a polarized molecule (24–26, 27, 29). Mo- lecular beam experiments have achieved high statistical sensitivity by measuring a large num- ber of molecules over a ≈1-ms coherence time (24, 25), whereas molecular ion–based exper- iments have used long Ramsey interrogation times (≈1 s), though with lower numbers (26, 27, 29). Measurements with trapped neu- tral polyatomic molecules can potentially com- bine the best features of each approach to achieve orders-of-magnitude improved statis- tical sensitivity (6). Experimental approach for polyatomic molecule control In this work, we demonstrate full quantum control over the internal states of a trapped polyatomic molecule in a vibrational bending mode with high polarizability in small electric fields. The protocol starts with preparing ultracold, optically trapped molecules in a single hyperfine level, after which a static elec- tric field is applied to polarize the molecules. The strength of the polarizing electric field is tuned to obtain near-zero g-factor spin states, which have strongly suppressed sensitivity to magnetic field noise while retaining eEDM sensitivity. Microwave pulses are applied to create a coherent superposition of these zero g-factor spin states, which precesses under the influence of an external magnetic field. The precession phase is then read out by a combi- nation of microwave pulses and optical cycling. We observed spin precession over a range of electric and magnetic fields and characterized the present limitations to the coherence time of the measurement. With readily attainable experimental parameters, coherence times on the order of the state lifetime (>100 ms) could be realistically achieved. We therefore realized the key components of an eEDM measure- ment in this system. Note that eEDM sensitiv- ity arises from the relativistic motion of the electron, which is enhanced by higher-mass nuclei. Although the light mass of CaOH pre- cludes a competitive eEDM measurement (30), the protocol demonstrated here is directly transferable to heavier laser-cooled alkaline earth monohydroxides with identical inter- nal level structures, such as SrOH, YbOH, and RaOH, which have substantially enhanced sen- sitivity to the eEDM (6, 11, 12, 30, 31). In eEDM measurements with polarized mol- → precesses under the ecules, the electron spin S influence of an external magnetic field BZ and the internal electric field of the molecule Eeff, which can be large owing to relativistic effects. Time evolution is described by the Hamiltonian H ¼ gSm → BBZ S (cid:2) ^Z (cid:3) deEeff S → (cid:2) ^n ¼ gSmBBZ MS (cid:3) deEeff S ð1Þ → → (cid:2) ^Z and S ¼ S Here, gS ≈ 2 is the electron spin g-factor, mB is the Bohr magneton, BZ points along the lab ^Z axis, and the internal field Eeff points along the mol- ecule’s internuclear axis ^n . We define the (cid:2) ^n to de- quantities MS ¼ S scribe the electron’s magnetic sensitivity and EDM sensitivity, respectively. The effect of the eEDM can be isolated by switching the orien- tation of the applied magnetic field or, alter- natively, by switching internal states to change the sign of MS or S. Performing both switches is a powerful technique for suppressing sys- tematic errors (25, 26). Present EDM bounds rely on specific states in diatomic molecules that have an unusually small g-factor, which reduces sensitivity to stray magnetic fields (24, 26). However, CaOH, like other laser-coolable molecules with structure amenable to eEDM searches (6, 31–33), has a single valence electron, which results in large magnetic g-factors. In this work, we engineered reduced magnetic sensitivity by using an ap- plied electric field EZ to tune MS to a zero- crossing while maintaining substantial eEDM sensitivity S. This technique is generic to poly- atomic molecules with parity doublets. Details of a specific M = ±1 pair of zero g-factor states are shown in Fig. 1, A and B, with further in- formation provided in (34). We emphasize that near the zero g-factor crossing, the eEDM sen- sitivity is nearly maximal because of the large projection of the electron spin onto the inter- nal electric field of the molecule. Sensitivity to transverse magnetic fields is also suppressed in these zero g-factor states (34). Experimental overview and single-state preparation The experiment begins with laser-cooled CaOH molecules loaded from a magneto-optical trap (14) into an optical dipole trap (ODT) formed by a 1064-nm laser beam with a 25-mm waist Anderegg et al., Science 382, 665–668 (2023) 10 November 2023 1 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Overview of the experiment. (A) A geometric picture of the bending → molecule at the zero g-factor crossing, showing that the electron spin (S ) has a finite → projection on the molecule axis (^n), giving eEDM sensitivity. However, the electron spin (S ) → ), which results in suppressed magnetic field is orthogonal to the magnetic field (B sensitivity. (B) The magnetic sensitivity (top) and eEDM sensitivity (bottom) for a pair of zero g-factor states (N = 1, J = 1/2+, F = 1, MF = ±1) are shown as a function of the applied electric field. (C) Experimental sequence to prepare the eEDM-sensitive state. First, the molecules are pumped into a single quantum state (N = 1, J = 1/2−, F = 0) with a combination of microwave drives and optical pumping (I). Next, a microwave p-pulse drives the molecules into the N = 2, J = 3/2−, F = 2, MF = 0 state (II). Lastly, the eEDM measurement state is prepared as a coherent superposition of the N = 1, J = 1/2−, F = 1, MF = ±1 states with a microwave p-pulse (III). The states that are optically detectable with the detection light are shown in black, whereas those not addressed by the detection light are in gray. → ¼ EZ ^Z and E size, as described in previous work (15). The ODT is linearly polarized, and its polarization ODT defines the ^Z axis, along which we vector D→ → ¼ also apply magnetic and electric fields, B ^Z , respectively, as depicted BZ in Fig. 1A. We first nondestructively image the molecules in the ODT for 10 ms as normaliza- tion against variation in the number of trapped molecules. The molecules are then optically pumped into the N = 1− levels of the ~X 2Sþ 010ð Þ vibrational bending mode (15) (Fig. 1C), and the trap depth is adiabatically lowered by 3.5 times to reduce the effect of ac Stark shifts from the trap light and to lower the temper- ature of the molecules to 34 mK. Any molecules that were not pumped into N = 1− levels of the bending mode are heated out of the trap with a pulse of resonant laser light. After transfer to the ~X Þ N ¼ 1(cid:3) Þ ð state, the molecular population is initially spread across 12 hyperfine Zeeman sublevels in the spin-rotation components J = 1/2 and 3/2. To prepare the molecules in a single hyperfine state, we used a combination of optical pump- ing and microwave pulses, as shown in Fig. 1C. We first applied microwaves from the N = 1, J = 3/2− state up to the N = 2, J = 3/2− state. Because this transition is parity-forbidden, we applied a small electric field EZ = 7.5 V/cm to slightly mix the parity of the N = 1 levels and provide transition strength. From the N = 2 state, we drove an optical transition to the ex- cited ~A Þ; J ¼ 1=2þ state. This state predominately decays to both the F = 0 2Sþ 010ð 2P 010ð Þk2S (cid:3)ð Fig. 2. Spin precession. (A) Spin precession of the eEDM-sensitive state in the presence of a bias magnetic field. Error bars represent 68% confidence intervals. au, arbitrary units. (B) Magnetic-field sensitivity of the eEDM state in CaOH as a function of electric field. The field sensitivity is determined by measuring the spin-precession frequency at different electric fields with an applied magnetic field of BZ = 110 mG. Error bars are smaller than the markers. The solid curve is the calculated magnetic-field sensitivity in the presence of trap shifts using known molecular parameters (34). (the target state) and F= 1 states in the N = 1, J = 1/2− manifold. After 3 ms of optical pump- ing, the microwaves were switched to drive the accumulated N = 1, J = 1/2−, F = 1 pop- ulation to the same N = 2, J = 3/2− state in ~X 010ð Þ, where they are excited by the optical light and pumped into the target F = 0 state. Once this optical pumping sequence is com- plete, we adiabatically ramped the electric field to EZ = 150 V/cm to substantially mix parity, then drove the population up to the N = 2, J = 3/2−, F = 2, M = 0 state with a microwave p-pulse (Fig. 1C, II). We cleaned out any remaining population in the N = 1 state with a depletion laser that resonantly drives the population to undetected rotational levels. Spin precession in an eEDM-sensitive state To perform spin precession in the eEDM- sensitive state, we first adiabatically ramped the electric field to a value EZ, then turned on a small bias magnetic field BZ. We measured the electron spin precession frequency using a procedure analogous to Ramsey spectroscopy (24, 25). The molecules were prepared by driving a p-pulse (2.5 ms), with microwaves linearly Anderegg et al., Science 382, 665–668 (2023) 10 November 2023 2 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Coherence time of the spin-precession signal. (A) Measured coherence times t versus BZ at different electric fields (red and blue markers, which correspond to different magnetic field sensitivities). The coherence time scales as 1/BZ owing to ac Stark-shift broadening then plateaus at a limit set by the magnetic field instability dB. This limit increases as the g-factor approaches zero. Solid and dashed curves are fit to the data. The ambient magnetic field noise determined from the fit is dB ¼ 4 fitted decoherence time due to light shifts is t ¼ 1=BZ 15 mG, near the zero g-factor crossing (0.06 MHz/G; red) and far from the crossing (1.0 MHz/G; blue). Shaded regions indicate the fitted exponential decay envelope of the oscillations; the 1.0 MHz/G data are excluded from the top panel for clarity. The spin-precession coherence time is extended by 16 times by approaching the zero g-factor point. þ20 (cid:3)10 ms (cid:4) mG. (B) Spin-precession signal, at BZ = þ2 (cid:3)1 mG, and the Þ (cid:4) 80 ð Fig. 4. Effect of trap light on coherence time. (A) Effective magnetic moment, meff, as a function of electric field for several trap intensities I. The trap light shifts the location of the zero crossing in meff. As a result, mole- cules at a finite temperature explore different magnetic-field sensitivities meff. (B) Dependence of the spin-precession frequency (scaled by the trap depth U0) on the position within the trap, where w0 is the trap waist. At lower magnetic fields, the relative change in spin- precession frequency is reduced. (C) Two spin-precession curves taken at the same magnetic field (BZ = 210 mG) but at different electric fields, showing that the ac Stark-shift limitation is independent of the effective g-factor because ac Stark shifts dominate the coherence time for large bias fields. ffiffiffi 2 jM ¼ (cid:3)1i polarized along the lab ^X axis, into the “bright” Þ= ð superposition state jBi ¼ M ¼ 1iþ j p within the N = 1, J = 1/2+, F = 1, M = ±1 eEDM-sensitive manifold (Fig. 1C, III). The state begins to oscillate between the bright state and the “dark” statejDi ¼ M ¼ 1i(cid:3) jM ¼ at a rate wSP= meffBZ, where the effec- (cid:3)1iÞ= tive magnetic moment m Bgeff ¼ gEm B eff Þ is tuned by means of ð (cid:3) hMSi hMSi the applied electric field EZ (Fig. 1B). The con- tribution from the deEeff term in Eq. 1 is neg- M¼(cid:3)1 ¼ m M¼1 ffiffiffi 2 p ð j ligible in CaOH but could be measured in heavier molecules with much larger Eeff. After a given time, a second p-pulse was applied to stop spin precession and transfer the bright state to the optically detectable N = 2, J = 3/2− level. Once the electric field was ramped down, the population remaining in the eEDM mani- fold, which has the opposite parity, is not op- tically detectable. We then imaged the ODT again and took the ratio of the first and second images (Fig. 2A). At long spin precession times (>10 ms), losses from background gas collisions (~1 s), blackbody excitation (~1 s), and the spon- taneous lifetime of the bending mode (~0.7 s) lead to an overall loss of signal, as characterized in (15). This effect is mitigated with a fixed dura- tion between the first and second images, making the loss independent of the precession time. To map out the location of the zero g-factor crossing, we performed spin precession mea- surements at a fixed magnetic field BZ = 110 mG for different electric fields (Fig. 2B). The spin precession frequency corresponds to an effec- tive g-factor at that electric field. We found that the zero g-factor crossing within the N = 1, J = 1/2+, F = 1, M = ±1 eEDM manifold occurs at an electric field of 59.6 V/cm, in agreement with theory calculations described in (34). We note that there is another zero g-factor cross- ing for the N = 1, J = 3/2+, F = 1 manifold at ≈64 V/cm, which has a smaller eEDM sen- sitivity but the opposite slope of geff versus EZ, thereby providing a powerful resource to re- ject systematic errors related to imperfect field reversals (34). We emphasize that although the location of these crossings is dependent on the structure of a specific molecule, their exist- ence is generic in polyatomic molecules, which naturally have parity-doublet structure (6). Coherence time and limitations A critical component of the spin precession measurement is the coherence time, which sets the sensitivity of an eEDM search. Figure 3A shows the measured coherence time of our system at different applied fields BZ and EZ. We characterized two dominant limitations that wash out oscillations at long times. Var- iations in the spin precession frequency can be linearly expanded as dwSP = meff(dBZ) + (dmeff)BZ. The first term describes magnetic field noise and drift of the applied bias field, given by dBZ. The second term describes noise and drifts in the g-factor, dgeff, which can arise from instability in the applied electric field, EZ, or from ac Stark shifts (described below). Drifts in the bias electric field EZ were found to be negligible in our apparatus. Decoherence caused by magnetic field noise, dBZ, is independent of the applied magnetic field but is proportional to meff and can be mitigated by operating near the zero g-factor crossing. As shown in Fig. 3B, at an electric field of 90 V/cm, corresponding to a large mag- netic moment of meff = 1.0 MHz/G, we realized a magnetic field noise–limited coherence time of 0.5 ms at BZ ≈ 15 mG (blue points). At an electric field of 61.5 V/cm, corresponding to meff = 0.06 MHz/G, which is much closer to the zero g-factor location, we found a coherence time of 4 ms at the same BZ (red points in Fig. 3B). At higher magnetic fields, the primary lim- itation to the coherence time is ac Stark shifts from the optical trapping light (Fig. 4). The intense Z-polarized ODT light leads to a shift Anderegg et al., Science 382, 665–668 (2023) 10 November 2023 3 of 4 RES EARCH | R E S E A R C H A R T I C L E in the electric field at which the zero g-factor crossing occurs. Owing to the finite tempera- ture of the molecules within the trap, they will explore different intensities of trap light and hence have different values of geff. The spread dgeff causes variation of wSP, which leads to de- coherence. In contrast to the magnetic field noise term, this effect is independent of the electric field EZ but decreases monotonically with BZ, which scales the frequency sensitivity to g-factor variations, dwSP = BZdmeff. The insen- sitivity of g-factor broadening to the exact value of geff is demonstrated in Fig. 4C. Decoher- ence caused by ac Stark shifts can be reduced by cooling the molecules to lower temperatures or by decreasing BZ. The bias magnetic field can be reduced arbitrarily far until either trans- verse magnetic fields or magnetic field noise becomes dominant. From the decoherence rates measured in this work, it is expected that ac Stark shift–limited coherence times of ~1 s could be achieved at bias fields of BZ ~ 100 mG. From the above discussion, it is expected that the longest achievable coherence times will occur for very small g-factors, geff ≈ 0, and very small bias fields, BZ ≈ 0. Minimizing BZ requires reducing the effects of both magnetic field noise and transverse magnetic fields to well below the level of the bias-field energy shifts. We canceled the transverse magnetic fields to below 1 mG by maximizing the spin precession period under the influence of trans- verse B fields only, and actively monitored and fed back on the magnetic field along each axis to minimize noise and drifts in BZ. Note that the stainless-steel vacuum chamber has no mag- netic shielding, which leads to high levels of magnetic field noise that would not be present in an apparatus designed for an eEDM search. Even under these conditions, we achieved a coherence time of 30 ms at an electric field of 60.3 V/cm (corresponding to meff = 0.02 MHz/G) and a bias field of BZ ≈ 2 mG (34). However, at such a low bias field, the molecules are sensi- tive to 60-Hz magnetic field noise that is pres- ent in the unshielded apparatus, which is on the same order as the bias field. Because the experiment is phase-stable with respect to the ac line frequency, this 60-Hz magnetic field fluctuation causes a time-dependent spin pre- cession frequency. Nevertheless, our prototype experiment confirms that long coherence times are possible. Any future eEDM experiment would have magnetic shielding that would greatly suppress nefarious magnetic fields from the environment. Such shielding could read- ily enable coherence times that exceed that of the ~0.5-s lifetime of the bending modes of similar linear polyatomic molecules with larger eEDM sensitivity (15). Discussion and outlook We have realized coherent control of optically trapped polyatomic molecules and demonstra- ted a realistic experimental roadmap for future eEDM measurements. By leveraging the dis- tinctive features of the quantum levels in poly- atomic molecules, we achieved a coherence time of 30 ms for paramagnetic molecules in a stainless-steel chamber with no magnetic shield- ing. With common shielding techniques used in past EDM experiments, there is a clear path to reducing stray fields and extending coher- ence times to >100 ms. At such a level, the dominant limitation becomes the finite lifetime of the bending mode (15). Even longer coher- ence times are possible with the right choice of parity-doublet states, as found in symmetric or asymmetric top molecules (6, 13, 35, 36). Following this roadmap with heavier trap- ped polyatomic molecules has the potential to provide orders-of-magnitude improvements to present bounds on T-violating physics. Using a recent study of the ~X 010ð Þ state in YbOH (37), we identified similar N = 1 zero g-factor states for eEDM measurements with greatly improved sensitivity. In addition to the g-factor tuning demonstrated in this work, polyatomic molecules provide the ability to reverse the sign of S without reversing MS, which is a crucial feature of recent experiments that has greatly improved the limit on the eEDM (25, 27). For example, in the N = 1 manifold of CaOH, there is another zero g-factor crossing at a nearby electric field value, with 69% smaller values of S and opposite sign. Because the ratio of eEDM sensitivity to g-factor versus EZ slope differs be- tween these two crossings, measurements at both points could be used to suppress systemat- ics caused by nonreversing fields that couple to the electric field dependence of the g-factor (25). This work provides an experimental dem- onstration of the advantages of the rich level structure of polyatomic molecules for preci- sion measurements. Although we focused here on spin precession with T-reversed states (M = ±1), many levels of interest can be favorably engineered for precision measurement exper- iments. In a recent proposal (9), parity doublets, magnetically tuned to degeneracy in optically trapped polyatomic molecules, were shown to be advantageous for searches for parity- violating physics. In another recent work (7), a microwave clock between rovibrational states in SrOH was proposed as a sensitive probe of ultralight dark matter, by using transitions tuned to electric and/or magnetic insensitiv- ity. In these proposals, and as now experi- mentally demonstrated in our work, coherent control and state engineering in polyatomic molecules can mitigate systematic errors and enable robust searches for new physics. RE FERENCES AND NOTES 1. M. L. Wall, K. Maeda, L. D. Carr, Ann. Phys. 525, 845–865 (2013). 2. M. Wall, K. Maeda, L. D. Carr, New J. Phys. 17, 025001 (2015). 3. Q. Wei, S. Kais, B. Friedrich, D. Herschbach, J. Chem. Phys. 135, 154102 (2011). 4. P. Yu, L. W. Cheuk, I. Kozyryev, J. M. Doyle, New J. Phys. 21, 093049 (2019). 5. L. D. Augustovičová, J. L. Bohn, New J. Phys. 21, 103022 6. 7. (2019). I. Kozyryev, N. R. Hutzler, Phys. Rev. Lett. 119, 133002 (2017). I. Kozyryev, Z. Lasner, J. M. Doyle, Phys. Rev. A 103, 043313 (2021). 8. N. R. Hutzler, Quantum Sci. Technol. 5, 044011 (2020). 9. E. B. Norrgard et al., Commun. Phys. 2, 77 (2019). 10. Y. Hao et al., Phys. Rev. A 102, 052828 (2020). 11. 12. B. L. Augenbraun et al., New J. Phys. 22, 022003 I. Kozyryev et al., Phys. Rev. Lett. 118, 173201 (2017). (2020). 13. D. Mitra et al., Science 369, 1366–1369 (2020). 14. N. B. Vilas et al., Nature 606, 70–74 (2022). 15. C. Hallas et al., Phys. Rev. Lett. 130, 153202 (2023). 16. M. Zeppenfeld et al., Nature 491, 570–573 (2012). 17. A. Prehn, M. Ibrügger, R. Glöckner, G. Rempe, M. Zeppenfeld, Phys. Rev. Lett. 116, 063005 (2016). 18. D. DeMille, J. M. Doyle, A. O. Sushkov, Science 357, 990–994 (2017). 19. T. E. Chupp, P. Fierlinger, M. J. Ramsey-Musolf, J. T. Singh, Rev. Mod. Phys. 91, 015001 (2019). 20. M. S. Safronova et al., Rev. Mod. Phys. 90, 025008 (2018). 21. C. Cesarotti, Q. Lu, Y. Nakai, A. Parikh, M. Reece, J. High Energy Phys. 2019, 59 (2019). 22. M. Pospelov, A. Ritz, Ann. Phys. 318, 119–169 (2005). 23. J. J. Hudson et al., Nature 473, 493–496 (2011). 24. J. Baron et al., Science 343, 269–272 (2014). 25. ACME Collaboration, Nature 562, 355–360 (2018). 26. W. B. Cairncross et al., Phys. Rev. Lett. 119, 153001 (2017). 27. T. S. Roussy et al., Science 381, 46–50 (2023). 28. R. Alarcon et al., arXiv:2203.08103 [hep-ph] (2022). 29. Y. Zhou et al., Phys. Rev. Lett. 124, 053201 (2020). 30. K. Gaul, R. Berger, Phys. Rev. A 101, 012508 (2020). 31. T. A. Isaev, A. V. Zaitsevskii, E. Eliav, J. Phys. At. Mol. Opt. Phys. 50, 225101 (2017). 32. T. A. Isaev, R. Berger, Phys. Rev. Lett. 116, 063006 (2016). 33. I. Kozyryev, L. Baum, K. Matsuda, J. M. Doyle, ChemPhysChem 17, 3641–3648 (2016). 34. See supplementary materials. 35. B. L. Augenbraun, J. M. Doyle, T. Zelevinsky, I. Kozyryev, Phys. Rev. X 10, 031022 (2020). 36. B. L. Augenbraun et al., Phys. Rev. A 103, 022814 (2021). 37. A. Jadbabaie et al., New J. Phys. 25, 073014 (2022). 38. L. Anderegg et al., Data for quantum control of trapped polyatomic molecules for eEDM searches. Zenodo (2023); https://doi.org/10.5281/zenodo.8384389. AC KNOWLED GME NTS A.J. acknowledges helpful discussions with C. Zhang and P. Yu. Funding: This work was supported by the Air Force Office of Scientific Research and the National Science Foundation (NSF). L.A. acknowledges support from the Harvard Quantum Initiative. N.B.V. from the Department of Defense National Defense Science and Engineering Graduate fellowship program, and P.R. from the NSF Graduate Research Fellowship Program. N.R.H. and A.J. acknowledge support from NSF CAREER program (PHY-1847550), the Gordon and Betty Moore Foundation (GBMF7947), and the Alfred P. Sloan Foundation (G-2019-12502). Author contributions: L.A., N.B.V., C.H., P.R., and A.J. performed the experiment and analyzed the data. N.R.H. and J.M.D. directed the study. All authors discussed the results and contributed to the manuscript. Competing interests: None declared. Data and materials availability: All data needed to evaluate the conclusions in this paper are present in the paper or in the supplementary materials. All data presented in this paper are deposited at Zenodo (38). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg8155 Supplementary Text Figs. S1 to S4 References (39–50) Submitted 23 January 2023; accepted 29 September 2023 10.1126/science.adg8155 Anderegg et al., Science 382, 665–668 (2023) 10 November 2023 4 of 4
10.1126_science.adg8116
RES EARCH MPOX APOBEC3 deaminase editing in mpox virus as evidence for sustained human transmission since at least 2016 Áine O’Toole1*, Richard A. Neher2, Nnaemeka Ndodo3, Vitor Borges4, Ben Gannon5, João Paulo Gomes4,6, Natalie Groves7, David J. King8, Daniel Maloney1, Philippe Lemey9, Kuiama Lewandowski5, Nicholas Loman7,10, Richard Myers7, Ifeanyi F. Omah1,11, Marc A. Suchard12, Michael Worobey13, Meera Chand7,14, Chikwe Ihekweazu3, David Ulaeto7†, Ifedayo Adetifa3†, Andrew Rambaut1*† Historically, mpox has been characterized as an endemic zoonotic disease that transmits through contact with the reservoir rodent host in West and Central Africa. However, in May 2022, human cases of mpox were detected spreading internationally beyond countries with known endemic reservoirs. When the first cases from 2022 were sequenced, they shared 42 nucleotide differences from the closest mpox virus (MPXV) previously sampled. Nearly all these mutations are characteristic of the action of APOBEC3 deaminases, host enzymes with antiviral function. Assuming APOBEC3 editing is characteristic of human MPXV infection, we developed a dual-process phylogenetic molecular clock that—inferring a rate of ~6 APOBEC3 mutations per year—estimates that MPXV has been circulating in humans since 2016. These observations of sustained MPXV transmission present a fundamental shift to the perceived paradigm of MPXV epidemiology as a zoonosis and highlight the need for revising public health messaging around MPXV as well as outbreak management and control. S ince 2017, the Nigeria Centre for Disease Control has been reporting cases of MPXV (mpox virus) infection in humans (fig. S1) (1). MPXV, a DNA virus in the genus Orthopoxvirus; Family Poxviridae, is often described as being endemic in West and Central Africa as a zoonotic disease that trans- mits through contact with the rodent reservoir host. Since the first human cases were observed in the 1970s, MPXV infections have been pre- dominantly associated with infants and children (2–4). However, of the cases observed in Nigeria since 2017, very few have been in children, with the virus mainly affecting adults aged 20 to 50 (79%), 27% of which were in women (5). Genome sequencing of viruses revealed enough genetic diversity among cases that distinct zoonotic events were not ruled out. In May 2022, cases 1Institute of Ecology and Evolution, University of Edinburgh, Edinburgh EH9 3FL, UK. 2Biozentrum, University of Basel and Swiss Institute of Bioinformatics, Basel, Switzerland. 3Nigeria Centers for Disease Control and Prevention, Abuja, Nigeria. 4National Institute of Health Doutor Ricardo Jorge (INSA), Lisbon, Portugal. 5UK Health Security Agency, Porton Down, Salisbury SP4 0JG, UK. 6Veterinary and Animal Research Centre (CECAV), Faculty of Veterinary Medicine, Lusófona University, Lisbon, Portugal. 7UK Health Security Agency, London E14 5EA, UK. 8CBR Division, Defence Science and Technology Laboratory, Salisbury SP4 0JQ, UK. 9Department of Microbiology, Immunology and Transplantation, Rega Institute, KU Leuven, Leuven, Belgium. 10University of Birmingham, Birmingham B15 2TT, UK. 11Department of Parasitology and Entomology, Nnamdi Azikiwe University, Awka, Anambra State, Nigeria. 12Department of Biostatistics, Fielding School of Public Health, University of California, Los Angeles, CA 90095, USA. 13Department of Ecology and Evolutionary Biology, University of Arizona, Tucson, AZ 85719, USA. 14UKHSA Guys and St Thomas’ NHS Trust, London SE1 7EH, UK. *Corresponding author. Email: aine.otoole@ed.ac.uk (Á.OT.); a.rambaut@ed.ac.uk (A.R.) †These authors contributed equally to this work. of MPXV infection were detected spreading widely across Europe and subsequently across the globe. The first MPXV genome sequences from these 2022 cases showed that they had descended from the clade characterized by cases diagnosed in Nigeria and Israel, Singapore, and the UK (6) associated with travel from Nigeria (fig. S2 and table S1, in bold). These early 2022 genomes are indicated as a triangle within clade IIb in Fig. 1A and represent lineage B.1 as per the nomenclature proposed by Happi et al. (7). Isidro et al. (6) noticed that sequences within lineage B.1 shared 42 single-nucleotide differences from the closest earlier MPXV ge- nomes from 2018. From a 2017 outbreak of MPXV in chimpanzees, the evolutionary rate of MPXV was estimated to be 1.9 × 10−6 substi- tutions per site per year (1.2 × 10−6 to 2.7 × 10−6), corresponding to ~1 nucleotide change every 3 years (8). Forty-two substitutions in the space of 3 to 4 years is an unexpectedly large number. Under the paradigm that MPXV is a zoonotic virus with limited human-to-human transmis- sion, one interpretation of this long branch might be that it represents adaptation to humans, facilitating the sustained transmission that is now observed. However, as we show here, and as was quickly seen when the first genomes from 2022 were sequenced, it is clear that these mutations are not the result of errors by the virus’s replication machinery and occur at a much higher rate than would be expected for an orthopoxvirus (6, 9). Specifically, most of observed nucleotide changes appear to be of a particular type—a dinucleotide change from TC→TT or its reverse complement, GA→AA (9, 10). This particular mutation is characteristic of the action of the APOBEC3 (apolipoprotein B mRNA editing enzyme, catalytic polypeptide 3) family of cytosine deaminases. These act on single-stranded DNA (ssDNA) to deaminate cytosine to uracil, causing a G→A mutation in the complementary strand when it is synthe- sized. Most human APOBEC3 molecules have a strong bias toward deaminating 5′TC dinu- cleotides, and APOBEC3-driven deamination has been demonstrated with many DNA viruses and retroviruses; (11–18). Furthermore, a recent study has specifically demonstrated APOBEC3F editing in cell culture and during human MPXV infection (19). We assess the extent to which APOBEC3 has acted on MPXV and explore whether this is the source of the elevated mutation rate observed since 2017. We also explore the evolutionary consequences of this mechanism driving evo- lution in MPXV and model the distinct pro- cesses underpinning the evolution of MPXV in the human population. APOBEC3 editing as a signature of MPXV evolution in the human population The known diversity of MPXV is decomposed into three major clades: clades I, IIa, and IIb (Fig. 1A) (7). Clade I represents MPXV sampled in Central Africa, and clade IIa is composed of viruses from human and nonhuman animal sam- ples taken in or connected to West Africa. Both of these clades include virus genomes spanning from the 1970s to the present day, although most samples were collected within the last 20 years (fig. S3). Clade IIb has an early sample taken in 1971 (GenBank accession KJ642617), but most of the sequences in clade IIb are more recent virus genomes from 2017 to 2022 that Happi et al. (7) have labeled as hMPXV-1 (Fig. 1A, labeled phy- logeny in fig. S2). Within the recent diversity of clade IIb (indicated as the darker box within IIb in Fig. 1A), we cataloged transmitted mutations that occurred between 2017 and 2022 with only a single representative from the 2022 global lineage B.1 (n = 44 genome sequences, fig. S2) (20). Within MPXV clade IIb, we observe rates of molecular evolution far greater than that ex- pected for double-stranded DNA viruses and indeed that observed in clades I and IIa of MPXV (8) and see this excess accumulation of mutation in samples as early as 2017. The great majority of these mutations are of the type G→A or C→T (90.8%) (Fig. 1B), regardless of sample host species (fig. S4). Comparing MPXV clade IIb with clade I and IIa emphasizes that this pattern is not seen outside of clade IIb (Fig. 1B, labeled phylogenies in fig. S5), nor is it seen when looking at reconstructed mutations within a phylogeny of 46 variola virus (VARV) genomes, the human virus responsible for smallpox (fig. S6). For the other MPXV clades, APOBEC-type mutations are observed at between 8 and 13% frequency, which fits with the expected proportion under standard models of nucleotide O’Toole et al., Science 382, 595–600 (2023) 3 November 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Specific enrichment of APOBEC3-type mutations in MPXV samples collected since 2017. (A) MPXV genetic diversity is categorized into clade I (predominantly sequences from the Democratic Republic of the Congo), clade IIa (predominantly West African sequences), and clade IIb. Within clade IIb is a subclade of genomes sampled from 2017 to 2022 that show mutational patterns distinct from those of the other two clades. (B) We catalog single-nucleotide mutations across the phylogenies of clade IIb, clade IIa, and clade I (top to bottom). For clade IIb, we include samples from 2017 to 2022 and only a single representative of the global lineage B.1. Of 120 reconstructed mutations that occurred on internal branches of the clade IIb phylogeny (so are observed transmitted mutations), 109 are consistent with APOBEC3 editing (90.8% of mutations). Individual proportions of G→A and C→T mutations are shown above the respective bars. Ancestral state reconstruction performed across clade IIa and clade I does not produce the same enrichment of mutations consistent with APOBEC3 editing, with only 27 of 207 observed mutations (13%) and 38 of 463 clade I mutations (8%) fitting the dinucleotide pattern. (C) Observed heptamers of C→T or G→A mutated sites of clade IIb, IIa, and I phylogenies (top to bottom). Heptamers associated with G→A mutations have been reverse-complemented to reflect deamination on the negative strand. For clade IIb, most C→T mutations are present in a TC dimer context, consistent with APOBEC3 editing (107 of 115 mutations, or 93%). However, the same is not seen for clades IIa and I, in which 29 of 149 (19%) mutations and 42 of 256 (16%) mutations have the dinucleotide context of APOBEC3, respectively, which is what we would predict under standard models of nucleotide evolution. *Only mutations occurring on internal branches of the clade IIb phylogeny are included. evolution (21, 22) (Fig. 1B). Notably, the hep- tamers of C→T and G→A mutations that oc- curred across the clade IIb phylogeny show that this is a specific enrichment of APOBEC3- type dimer mutations of the type TC→TT and GA→AA (Fig. 1C). Similarly, this enrichment of TC→TT and GA→AA mutations is observed within the B.1 lineage (fig. S7), where 84.8% of observed single-nucleotide polymorphisms are consistent with APOBEC3 editing (fig. S8). Observed heptamers around the observed C→T and G→A mutations show a pronounced en- richment in TC and GA target sites in the ge- nomes sampled from 2017 to 2022, in contrast with the rest of MPXV diversity (Fig. 1C), and this enrichment is also reflected within line- age B.1 (fig. S9). Our analysis highlights that evolution within clade IIb before the emergence of lineage B.1 mirrors that within lineage B.1 but is distinct from MPXV clade I or IIa. Since 2022, the B.1 lineage has been sampled and sequenced inter- nationally in the global epidemic of MPXV. Lineage B.1 is known to be circulating by sus- tained human-to-human transmission and as such, mutations that have accumulated in B.1 can be considered characteristic of this mode of transmission. We suggest that the APOBEC3- driven evolution of recent clade IIb MPXV is a signature of a switch to sustained transmission within the human population. Within the B.1 lineage, believed to be entirely the result of hu- man infection and transmission, we continue to see the same pattern of predominantly APOBEC3 mutations accumulating at a rate similar to that seen in A lineage genomes since 2017. It is unlikely that, by chance, MPXV evolved to become susceptible to APOBEC3 action within the putative rodent reservoir before the emer- gence of cases and to retain that susceptibility to human APOBEC3 molecules once transmitting in humans. Given that all human cases se- quenced since 2017 share substantial numbers of APOBEC3 mutations, including nine on the stem branch leading to hMPXV-1, it is very un- likely that these represent multiple zoonotic introductions. APOBEC3 genes emerged in placental mammals from a duplication of the ancestral AID gene and have a dynamic recent evolutionary past, with gene duplication and loss across phyla (23, 24). APOBEC3 genes in primates have undergone recent expansion, with primate genomes now having seven paralogs of three ancestral genes (25, 26). Rodents, the reservoir of MPXV, have only a single functional APOBEC3 protein, resulting from gene loss and fusion events (25). Rodent APOBEC3 has been shown to be expressed preferentially in spleen and bone marrow, with limited expression ob- served in other tissues (27, 28). O’Toole et al., Science 382, 595–600 (2023) 3 November 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Observed APOBEC3-type mutations are not merely a product of available target sites. (A) Consequence of hypothetical APOBEC3 mutations at target dimer site (either the C in the TC target site or G in the GA target site) in the coding regions of the National Center for Biotechnology Information reference MPXV genome for clade II (accession NC_063383) and those observed APOBEC3 mutations across the coding regions of the clade IIb phylogeny (not including the outgroup branch leading to the 1971 genome sequence). These are categorized into nonsynonymous (altered amino acid), synonymous (amino acid remaining unchanged), nonsense (editing producing a stop codon), and intergenic (not present in a coding sequence). (B) The proportion of target sites edited for each target site percentile window across the MPXV genome. The teal shaded regions represent the binomial confidence interval around observations. Masked regions are indicated by vertical gray bars. Observed edits include data from the clade IIb phylogeny, with a single representative of lineage B.1 and not including the branch leading to the outgroup 1971 genome sequence. (C) Hypothetical amino acid changes for codons overlapping with TC and GA target sites in a reference MPXV genome (GenBank accession number: NC_063383) if APOBEC3 edited those dimers to TT and AA. Amino acid changes are colored by Grantham Score (0–50 conservative, dark blue; 51–100 moderately conservative, light blue; 101–150 moderately radical, light red; >150 radical, dark red; synonymous, gray). Single-letter abbreviations for the amino acid residues are as follows: C, Cys; D, Asp; E, Glu; F, Phe; G, Gly; H, His; I, Ile; K, Lys; L, Leu; M, Met; N, Asn; P, Pro; Q, Gln; R, Arg; S, Ser; W, Trp; and Y, Tyr. Stop codons are indicated with an asterisk (*). APOBEC3 has a limited repertoire to generate variation in the MPXV genome If we assume this observed evolution within hMPXV-1 is APOBEC3-driven, this may have implications for its sustained transmission in the human population. Considering all GA and TC dimer sites in the clade II reference ge- nome (accession number NC_063383)—i.e., those that could be the target of APOBEC3 editing but had not been by that point—we assessed what amino acid changes a deamination mutation at these sites would bring about (Fig. 2). Of the 23,718 such dimers, 61.6% (14,618) would pro- duce amino acid replacements, 21% (4990) would be synonymous, 2.9% (692) would in- duce stop codons, and 14.4% (3418) would occur outside of coding regions. For the clade IIb ge- nomes, of the 633 mutations at these dimers that did occur, 38.7% (245) were amino acid replacements and 35.7% (226) were synony- mous, 4.7% (30) were nonsense, and a further 132 APOBEC3 mutations were in intergenic regions (20.8%). The probability of getting 226 or greater synonymous mutations out of 663 under a simple binomial distribution with 21.0% chance of a context being synonymous is P = 7.6 × 10−18. We do not see the same enrichment for synonymous mutations in the mutations that are not APOBEC3-like, although the quantity of these mutations is considerably lower (fig. S10). O’Toole et al., Science 382, 595–600 (2023) 3 November 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E There are also more mutations outside of protein- coding regions than we would expect based on the location of target dimers (probability of 4.5 × 10−6 of getting at least 132 noncoding muta- tions given that only 14.4% of targets are in these regions). This supports the hypothesis that what we observe are the residual least harmful APOBEC3 mutations after natural selection has eliminated those with substan- tial fitness costs to the virus. By comparing the density of observed C→T and G→A mutations and the density of APOBEC3 target sites (TC or GA dinucleotides) across the reference ge- nome, we see that the distribution of muta- tions is not simply a product of the availability Fig. 3. Estimating the time of MPXV emergence into the human population from the accumulation of APOBEC3-type mutations. (A) MPXV genomes sampled from human infections from 2017 to 2022, with an outgroup sequence from an outbreak in Nigeria in 1971 (n = 44 genome sequences, including outgroup). Lineages indicated as per nomenclature proposed by Happi et al. (7). Mutations along each branch are indicated with circles colored by whether it is putatively APOBEC3 edited (TC→TT and GA→AA; red) or whether it is another mutation type (yellow). The break in the basal branch illustrates the assumption made in the regression model in panel B, that all APOBEC3 mutations occurred after emerging into the human population; however, we do not know the precise distribution of red or yellow mutations. (B) APOBEC3 mutations from MPXV genomes sampled since 2017. The reconstructed most recent common ancestor (MRCA) of the panel A phylogeny is used as the root in the root-to-tip plot, and the y intercept is used as a proxy for time of emergence, which is inferred by fitting a Bayesian regression to the sequence dates from panel A. Intersects with y = 1, 2, and 3 are also shown as it is likely that a small number of the APOBEC3-type mutations are actually earlier replication errors and not induced by APOBEC3. (C) Maximum clade credibility (MCC) phylogeny of MPXV clade IIb with absolute time shown on the x axis. We separated the alignment into an APOBEC3 and a non-APOBEC3 partition and modeled the substitution process in each independently. We used an epoch model with two outgroup sequences (not shown) representing the first epoch and hMPXV-1 ingroup sequences representing MPXV post-emergence into the human population with an exponential growth model. The probability density distributions show the estimated time of the most recent common ancestor (tMRCA) of the ingroup as well as the estimated transition time that represents the time of emergence into the human population. (D) Estimated effective population size of the outbreak using a nonparametric coalescent Skygrid model with 11 change points over a period of 8.5 years. This reconstruction falls within the bounds of the exponential growth model estimated from the second epoch in panel C, suggesting that the MPXV population has been exponentially growing since at least 2016. O’Toole et al., Science 382, 595–600 (2023) 3 November 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E of target sites (Kolmogorov-Smirnov test sta- tistic = 0.07, P = 0.0002; fig. S11, A and B). When considering synonymous and nonsynonymous APOBEC3-like mutations separately, there is a significant difference between the density of target sites across the MPXV genome and the distribution of observed APOBEC3-like synon- ymous and nonsynonymous mutations, respec- tively (fig. S11, C and D). The “repertoire” of mutations that APOBEC3 is able to provide as genetic variation on which natural selection can act is severely restricted. Only a limited number of dinucleotide contexts are present, and the repertoire of amino acid changes that APOBEC3 editing can induce is also limited (Fig. 2C and fig. S12). Only 13 dif- ferent amino acid replacements are possible, and three that give rise to stop codons, and they are not reversible by the same mechanism. This means that given the restricted set of positions at which these mutations occur and the limited amino acid changes that they can result in, the elevated rate will not necessarily facilitate adap- tation of the virus. APOBEC3 hypermutation is a host-mediated antiviral mechanism. These molecules act as the viral genome is being replicated and single strands are exposed. During repeated rounds of replication, either strand can be deaminated, leading to both C→T and G→A changes on the positive strand, as seen in this study. Thus, it is likely that the genomes that are extensively mu- tated by APOBEC3 will simply not be viable and will not be transmitted further. MPXV replicates in the cytoplasm likely by means of rolling-circle amplification (29), and this facilitates exten- sive continuous genome replication (30). The high processivity of this mechanism efficiently produces high copy numbers of MPXV genome molecules in the cell, potentially saturating the APOBEC3 enzyme action if the concentration of MPXV DNA molecules is high enough. This likely means that many MPXV genomes are un- affected by APOBEC3 action. Occasionally, how- ever, a genome, modestly mutated by APOBEC3, may remain viable and be transmitted. We see this in the enrichment of observed synonymous and intergenic mutations relative to available targets in the MPXV genome. Given the non- reversibility of the APOBEC3 action, sustained evolution within the human population may result in a depletion of lower-consequence target sites (i.e., synonymous or conservative amino acid changes) and thus expedite a decrease in fitness of MPXV over time. This could be both through a reduction in the number of viable offspring viruses produced by infected cells and as a result of the accumulation of moderately deleterious mutations by genetic drift (i.e., mu- tation load). However, the timescale on which this might happen is uncertain and other evolu- tionary forces such as recombination may act to restore fitness, and we do not address this fur- ther in this study. A further uncertainty arises when considering the variable repertoire of genes associated with virus infectivity or host immune modulation in poxvirus genomes. Mu- tations that alter or abrogate the function of these genes may have little direct effect on virus replication machinery but may disrupt the virus–host interaction. There is precedent for the naturally occurring inactivation of genes in VARV contributing to host specificity, and consequently the loss of function of some MPXV genes through APOBEC3 activity may poten- tially have adaptive value for the virus as it repli- cates and transmits in a new host (31, 32). Even if the mutations that accumulate through this process are simply the neutral residue of a suboptimal antiviral host defense, they have produced sufficient variability for the phyloge- netic analysis of the epidemic over the short term. The initial lineages proposed by Happi et al. (7) have expanded with the 2022 epidemic B.1 lineage now encompassing 17 sublineages at the time of writing (33). The rapid and temporally linear accumulation of mutations means that genomic epidemiological models and tools (34, 35), usually used for RNA viruses, may also have utility for hMPXV-1. The linear accumulation of APOBEC3-type mutations since the emergence and spread of MPXV in humans Since 2017, the genomes thus far sampled from clade IIb have accumulated APOBEC3-type single-nucleotide mutations approximately linearly over time (Fig. 3, A and B; labeled phylog- eny in fig. S13). We applied Bayesian regression analysis on the root-to-tip plot of sequences in Fig. 3A, which includes one representative B.1 genome, and also separately on the B.1 lineage (B.1 phylogeny in fig. S1) (20). To ensure that the elevated temporal signal is specific to APOBEC3 data, we show combinations of APOBEC3 and non-APOBEC3 mutations on clade IIb data in fig. S14. The estimated rate of accumulation was 6.18 per year (95% credible intervals of 5.20, 7.16). For the B.1 lineage, the rate was 5.93 per year (2.95, 8.92), suggesting that despite widespread and rapid transmission within MSM (men who have sex with men) networks, the rate of accumulation of APOBEC3 mutations re- mained the same as it did for the rest of clade IIb. It is notable that the regression line for B.1 lies substantially above that for the rest of clade IIb, suggesting that this lineage accumulated more mutations than expected before the emergence of B.1. However most of these mu- tations are also present in the genome from Maryland, USA (accession number ON676708) from November 2021 (10), indicating that they arose and circulated for some months before the B.1 epidemic (figs. S15 and S16). Extrapolat- ing back to when the APOBEC3-type mutations started to accumulate provides an estimate of when the first APOBEC3 mutations occurred in the stem of the branch leading to the 2017 epidemic. If we assume that all these mutations are due to APOBEC3 in humans, then we esti- mate this date of emergence to be 20 June 2015. However, we expect a few APOBEC3-like muta- tions to actually be due to replication errors during the earlier preemergence epoch. The num- ber of mutations that we ascribe to this period will affect our estimate linearly—i.e., if three mutations were not due to APOBEC3, then the estimate would shift to 14 December 2015. To accommodate this uncertainty in our es- timates, we have developed a more explicit model of APOBEC3-mediated evolution in the BEAST software package (20, 34). We estimate that the action of APOBEC3 on the MPXV pop- ulation is driving evolution ~28 times faster than the background evolutionary rate (fig. S17). Gigante et al. (10) also described an elevated overall rate of evolution in the A lineage but did not decompose the APOBEC3 and non- APOBEC3 contribution to this. The time of the most recent common ancestor of the post 2017 genomes is estimated to be 23 February 2016 (28 June 2015, 28 September 2016), with the transition to sustained human-to-human trans- mission estimated to be 14 September 2015 (21 August 2014, 31 July 2016; Fig. 3C and fig. S17). Unlike the assumption in Fig. 3B that all APOBEC3 mutations occurred after emergence, the BEAST analysis estimates the transition point integrating over all possibilities. This allows for the fact that a few of the APOBEC3- like mutations may actually be due to replica- tion errors in the earlier evolutionary epoch and this might explain the slightly more recent date. We also see evidence of exponential growth in the number of infections in the epi- demic before the emergence of lineage B.1 in 2022 (Fig. 3D), despite the decline in cases reported in 2020 (fig. S1C), albeit the growth rate is relatively slow, which reinforces the indi- cation from the demographics of the cases that this is not a generalized epidemic. Implications for the global public health response to mpox cases Since the identification of the B.1 lineage, a number of countries have reported other line- ages that lie outside the diversity of B.1, in- cluding the United States, United Kingdom, Portugal, India, and Thailand. In almost all instances, these cases are reported as having a history of international travel. The lineages in which these are placed (designated as A.2.1, A.2.2, A.2.3, and A.3) can all be phylogenet- ically traced back to the epidemic in Nigeria (Fig. 3A). This suggests that at least one in- stance of sustained human-to-human transmis- sion is still ongoing outside of the recognized MSM networks that were the focus of the 2022 global epidemic. Stopping transmission in these communities, though necessary, will not be sufficient to eliminate the virus as a human epidemic. Many countries lack the surveillance O’Toole et al., Science 382, 595–600 (2023) 3 November 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E to detect MPXV cases, and if sustained human- to-human transmission has been ongoing since 2015–2016, it is plausible that there are other populations that are currently enduring epidemics. Historically, mpox was considered a zoonotic disease, and cases have been treated as inde- pendent spillover events with low levels of cir- culation in the human population. Thus far, this continues to be an accurate characteriza- tion of clade I in Central Africa. For clade IIb, although some non-B.1 lineage cases may be new zoonotic infections, most cases since 2016 are likely the result of human-to-human trans- mission. Although the B.1 lineage across the world is now diminished—though not yet eradicated—the human epidemic from which it arose continues unabated. It is critical that global public health affords MPXV cases in countries that are historically considered to have endemic reservoir species equal attention and concern to those elsewhere. Surveillance needs to be global if MPXV is to be eliminated from the human population and then prevented from reemerging. RE FE RENCES AND N OT ES 1. NCDC, “An Update of Monkeypox Outbreak in Nigeria” (Nigeria Centre for Disease Control, 2017), (available at https://ncdc. gov.ng/diseases/sitreps/?cat=8). 2. A. W. Rimoin et al., Proc. Natl. Acad. Sci. U.S.A. 107, 16262–16267 (2010). 3. T. D. Baker, in Clio Medica. Acta Academiae Internationalis Historiae Medicinae (Brill, 1982), vol. 17, pp. 268–269. 4. Z. Jezek et al., J. Trop. Med. Hyg. 90, 31–38 (1987). 5. E. Alakunle, U. Moens, G. Nchinda, M. I. Okeke, Viruses 12, 1257 (2020). J. Isidro et al., Nat. Med. 28, 1569–1572 (2022). 6. 7. C. Happi et al., PLOS Biol. 20, e3001769 (2022). 8. L. V. Patrono et al., Nat. Microbiol. 5, 955–965 (2020). 9. Á. O’Toole, A Rambaut, Initial observations about putative APOBEC3 deaminase editing driving short-term evolution of MPXV since 2017. Virological (2022); https://virological.org/t/830. 10. C. M. Gigante et al., Science 378, 560–565 (2022). 11. R. Suspène, M. Henry, S. Guillot, S. Wain-Hobson, J.-P. Vartanian, J. Gen. Virol. 86, 125–129 (2005). 12. H. Chen et al., Curr. Biol. 16, 480–485 (2006). 13. J.-P. Vartanian, D. Guétard, M. Henry, S. Wain-Hobson, Science 320, 230–233 (2008). 14. J.-P. Vartanian et al., PLOS Pathog. 6, e1000928 (2010). 15. M. Tsuge et al., Virus Res. 149, 211–216 (2010). 16. R. Suspène et al., Proc. Natl. Acad. Sci. U.S.A. 108, 4858–4863 (2011). 17. Z.-G. Peng et al., Hepatology 53, 1080–1089 (2011). 18. M. Fehrholz et al., J. Gen. Virol. 93, 565–576 (2012). 19. R. Suspène et al., J. Infect. Dis.jiad165 (2023). 20. See materials and methods. 21. M. Hasegawa, H. Kishino, T. Yano, J. Mol. Evol. 22, 160–174 (1985). 22. M. Kimura, J. Mol. Evol. 16, 111–120 (1980). 23. S. G. Conticello, C. J. F. Thomas, S. K. Petersen-Mahrt, M. S. Neuberger, Mol. Biol. Evol. 22, 367–377 (2005). 24. A. Krishnan, L. M. Iyer, S. J. Holland, T. Boehm, L. Aravind, Proc. Natl. Acad. Sci. U.S.A. 115, E3201–E3210 (2018). 25. C. Münk, A. Willemsen, I. G. Bravo, BMC Evol. Biol. 12, 71 (2012). 26. L. Yang, M. Emerman, H. S. Malik, R. N. McLaughlin Jnr., eLife 9, e58436 (2020). 27. M. C. Mikl et al., Mol. Cell. Biol. 25, 7270–7277 (2005). 28. A. I. Su et al., Proc. Natl. Acad. Sci. U.S.A. 99, 4465–4470 (2002). 29. B. Moss, Cold Spring Harb. Perspect. Biol. 5, a010199 (2013). 30. Q. Peng et al., Science 379, 100–105 (2023). 31. R. C. Hendrickson, C. Wang, E. L. Hatcher, E. J. Lefkowitz, Viruses 2, 1933–1967 (2010). 32. B. Mühlemann et al., Science 369, eaaw8977 (2020). 33. GitHub, mpxv-lineages/lineage-designation: Official place for proposals and details around monkeypox virus lineage designations; https://github.com/mpxv-lineages/lineage-designation. 34. M. A. Suchard et al., Virus Evol. 4, vey016 (2018). 35. P. Sagulenko, V. Puller, R. A. Neher, Virus Evol. 4, vex042 (2018). 36. A. Rambaut, Á. O’Toole, hmpxv/apobec3, Zenodo (2023); https://doi.org/10.5281/zenodo.8234483. ACKN OWLED GMEN TS We thank those involved in sequencing MPXV genome sequences and shared data online rapidly, facilitating rapid public health responses, and also those who have continued to contribute to open science by sharing their data in open access platforms such as GenBank. We have provided all data and scripts related to this manuscript on Zenodo (36). Funding: Wellcome Trust ARTIC (Collaborators Award 206298/Z/17/Z, ARTIC network) (Á.O.T., P.L., M.A.S., A.R.); European Research Council (grant agreement no. 725422 – ReservoirDOCS) (P.L., M.A.S., A.R.); National Institutes of Health (R01 AI153044) (P.L., M.A.S., A.R.); David and Lucile Packard Foundation (M.W.); Research Foundation, Flanders– Fonds voor Wetenschappelijk Onderzoek–Vlaanderen, G066215N, G0D5117N and G0B9317N (P.L.); HORIZON 2020 EU grant 874850 MOOD (P.L.); HERA project (grant/2021/PHF/23776) supported by the European Commission through the European Centre for Disease Control and Prevention (V.B. and J.P.G.). Crown Copyright © 2023. The Nigeria Centre for Disease Control and Prevention receives core funding from the Nigerian government. Author contributions: Conceptualization: Á.O.T., A.R.; Methodology: Á.O.T., R.N., A.R.; Investigation: N.N., V.B., B.G., J.P.G., N.G., D.K., D.M., K.L., R.M., I.F.O., M.W.; Visualization: Á.O.T., R.N., A.R.; Funding acquisition: A.R., P.L., M.S., M.C., C.I., D.U., I.A.; Supervision: D.U., I.A., A.R.; Writing–original draft: Á.O.T., R.N., P.L., M.S., M.W., A.R.; Writing– review and editing: Á.O.T., R.N., N.N., V.B., B.G., J.P.G., N.G., D.K., D.M., P.L., K.L., N.L., R.M., I.F.O., M.S., M.W., M.C., C.I., D.U., I.A., A.R. Competing interests: The authors declare that they have no competing interests. Data and materials availability: Data, code, and materials used in the analysis are available from Zenodo (36). Tables S1 and S2 include details of accession numbers and author lists for all source genome sequence data used. All except a single genome sequence were sourced from GenBank on NCBI. Data partitions of the alignments have been included in the XML files on Zenodo and GitHub for ease of reproduction. A single sequence was sourced from GISAID (gisaid.org) and as such was removed from the XMLs. Supplementary tables include the GISAID identifier, which can be used to independently source the data, and instructions for constructing the alignment and addition to the XML are included. License information: Copyright © 2023 the authors, some rights reserved, including, where applicable, UK Crown copyright licensed under the Open Government License v.3.0; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse. In the interest of rapid dissemination of results with immediate public health relevance, and because this research was funded in whole or in part by the Wellcome Trust (Collaborators Award 206298/Z/17/Z, ARTIC network), a cOAlition S organization, the author will make the Author Accepted Manuscript (AAM) version available under a CC BY public copyright license. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg8116 Materials and Methods Figs. S1 to S17 Tables S1 to S4 References (37–40) MDAR Reproducibility Checklist Submitted 23 January 2023; resubmitted 1 June 2023 Accepted 21 September 2023 10.1126/science.adg8116 O’Toole et al., Science 382, 595–600 (2023) 3 November 2023 6 of 6
10.1126_science.adg2657
RES EARCH CHIRAL MATERIALS A magnetic assembly approach to chiral superstructures Zhiwei Li†, Qingsong Fan, Zuyang Ye, Chaolumen Wu, Zhongxiang Wang, Yadong Yin* Colloidal assembly into chiral superstructures is usually accomplished with templating or lithographic patterning methods that are only applicable to materials with specific compositions and morphologies over narrow size ranges. Here, chiral superstructures can be rapidly formed by magnetically assembling materials of any chemical compositions at all scales, from molecules to nano- and microstructures. We show that a quadrupole field chirality is generated by permanent magnets caused by consistent field rotation in space. Applying the chiral field to magnetic nanoparticles produces long-range chiral superstructures controlled by field strength at the samples and orientation of the magnets. Transferring the chirality to any achiral molecules is enabled by incorporating guest molecules such as metals, polymers, oxides, semiconductors, dyes, and fluorophores into the magnetic nanostructures. S uperstructures of colloidal particles can assemble with chiral symmetry (1–5). The driving force for chiral assembly is that the particles or structures are intrin- sically chiral or rendered so with ad- sorbed surface molecules and templates that create chirality (usually helical structures) or by lithographic methods (6–8). Chiral super- structures, especially those made from plas- monic nanoparticles, have distinctive optical properties such as circular dichroism (CD) under circularly polarized light excitation (9–14) and have the potential for developing electric and optical sensors and devices (15–20). Templated assembly and lithography have been used to create chiral superstructures for sensing external stimuli through changes in CD spectra (21, 22). For example, DNA- templated assembly can transfer the helical configuration of DNA templates to many nano- structures (23–26) and be used to monitor changes in temperature and chemical bind- ing (27–31). Controlling the collective orientation of these chiral structures in either solution or solid matrices remains challenging for optimiz- ing chiroptical performance. Additionally, the existing strategies such as templated assem- blies and chiral superlattice formation only work for materials of narrow length scales and specific chemical compositions or shapes (32–34). Designing miniature chiroptical de- vices and understanding light-matter interac- tions that involve distinct physical principles would benefit from a general approach for assembling achiral materials of diverse sizes, shapes, and chemical compositions into chiral superstructures with actively tunable chiroptical responses (35–37). Department of Chemistry, University of California, Riverside, CA 92521, USA. *Corresponding author. Email: yadong.yin@ucr.edu †Present address: Department of Chemistry and International Institute for Nanotechnology, Northwestern University, Evanston, IL 60208, USA. Here, we report the rapid and reversible assembly of materials of varying compositions and length scales, from small molecules to nano- and microstructures, into chiral struc- tures, and the active tuning of the structural handedness, collective orientation, and chirop- tical properties by using the magnetic field of a permanent magnet. Our analytical model demonstrates the presence of a quadrupole field chirality in the gradient magnetic field of the magnet. Assembling nanorods in such a magnetic field led to the formation of chiral superstructures, with their handedness and chiroptical properties being determined by magnet position and orientation. The struc- tural chirality could be transferred to organic molecules and inorganic compounds by dop- ing into or coating onto the host magnetic building blocks, demonstrating a general approach to chiral superstructures. The quadrupole field chirality of permanent magnets We consider the magnetic field and field dis- tribution of a cube-shaped permanent mag- net (fig. S1), with the north pole of the magnet pointing upward in fig. S1A. The calculated azimuth of the local magnetic field (mapped in Fig. 1A) undergoes gradual changes in the field direction in each quadrant. A differential magnetic field was calculated by subtracting magnetic field vectors in two chosen yz cross sections (fig. S2). In Fig. 1B, the resulting local- field rotation is indicated by the black arrows, and the magnitude is color mapped, with the rotation angle (Dw) being defined as the dif- ferences between the azimuth of the magnetic fields in the two cross sections. The differential field forms a quadrupole, with two left-handed (positive Dw in red domains) and right-handed (negative Dw in blue domains) magnetic field domains. Figure S3 further delineates the rotation of local fields along a chosen path- way, demonstrating the helical magnetic field distribution. We developed an analytical model to under- stand the assembly of magnetic nanorods in such a chiral field (supplementary materials) that could predict magnetic nanorod align- ment along the local field to form chiral superstructures (Fig. 1C and fig. S4). For small magnetic nanorods (107.6 ± 5.2 nm in length and 13.0 ± 1.7 nm in diameter), there were no obvious CD signals in rod dispersion without a magnetic field or in a magnetic field along the x axis (Bx), but changing the field direc- tion from the x axis to the y and z axes cre- ated CD responses (Fig. 1, D and E). We observed positive and negative CD peaks of similar intensity for y (By) and z fields (Bz), respectively, which suggested chiral super- structures with opposite handedness. The CD responses were then measured in an aqueous solution of glycerol (n = 1.475) with an in- creasing volume ratio from 0 to 100%. The increase in effective refractive index (n) of the solution induced consistent CD-intensity decrease under the same magnetic fields (fig. S5). The nanorods’ CD signal weakened but did not disappear as the solution refractive index approached that of the silicon dioxide (SiO2) layers (n = 1.475 ± 0.005) (38). This refractive index–matching experiment demon- strates that both the scattering and absorption of the nanorods contribute to the overall CD responses. We calculated an optical anisotropic factor (g factor) of ~0.01 at 400 nm (fig. S6). To verify the formation of chiral superstructures, cyanine 3–doped Fe3O4@SiO2 core-shell nano- rods (322.2 ± 16.5 nm in length, 70.2 ± 4.7 nm in diameter, and 50.3 ± 1.5 nm in silica thickness) were used as magnetic building blocks (39, 40) and fixed in a polymer by photocuring under a uniform magnetic field. Linear chains were formed because of magnetic dipole-dipole interactions and were parallel to the uniform field with a standard deviation of 0.36° to minimize the demagnetizing fields (figs. S7 and S8) (41). If a gradient field of a perma- nent magnet (cube shape, 12 mm in edge length) was used (fig. S9), the chain alignment within one yz plane and the chain rotation between different yz planes were determined by the local magnetic fields and field rotation, respectively (figs. S10 and S11). Thus, the chiral superstructures made of large nanorods show similar CD responses to magnetic fields (fig. S12). Additionally, we characterized the chain alignment in 10 sequential yz layers from x = 1 to 10 mm and in five layers along the y axis using optical and electron microscopy, confirming the chain rotation into chiral superstructures as driven by the quadrupole chiral field (figs. S13 to S17). Magnetic assembly and active tuning of plasmonic chiral superstructures To systematically study the CD dependence on magnetic fields over a wide spectral range, Li et al., Science 380, 1384–1390 (2023) 30 June 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. The quadrupole field chirality of a permanent magnet. (A) Normalized magnetic field (white arrows) and field-vector azimuth (color maps) of a permanent magnet (high- lighted in the middle square) with a cubic shape and edge length of 2 cm. (B) The field rotating vectors (black arrows) and field angle changes (color map) of the magnetic field along the x axis. Positive rotation angles (Dw) represent clockwise left-handed rotation of the magnetic field, and negative rotation angles represent counterclockwise right-handed field rotation. deg, degree. (C) Schematic illustra- tion of the magnetic assembly during CD measurement and simulated helical super- structures assembled from magnetic nanorods under such a chiral magnetic field. (D) Extinction spectrum and (E) CD spectra of Fe3O4@SiO2 nanorods under different magnetic fields. Arbitrary units, a.u.; mdeg, millidegree. A 0 ) m c ( Z -1 -2 -3 90 -90 -4 -3 -2 -1 1 2 3 0 Y (cm) C z x 0o 180o y k x (0,1,-0.75) z x z y B 0 ) m c ( Z −2 (deg) 6.10 −4 −3 −2 −1 0 1 2 3 −6.10 Y (cm) D E n o i t c n i t x E ) . . u a ( 3 2 1 0 300 ) g e d m ( D C 900 0 -900 300 450 600 Wavelength (nm) 750 900 no magnet By 600 Wavelength (nm) 450 750 Bx Bz 900 we introduced Fe3O4/Au hybrid nanorods as building blocks by taking advantage of the localized surface plasmon resonance (LSPR) of Au nanorods and the magnetic responses of Fe3O4 nanorods. The Au nanorods were synthesized with a space-confined growth method and had a length of 156.6 ± 15.2 nm and a diameter of 48.9 ± 4.7 nm (42, 43). As shown in fig. S18A, Fe3O4@SiO2 nanorods were introduced (107.6 ± 5.2 nm in length, 13.0 ± 1.7 nm in diameter, and 5.0 ± 0.5 nm in silica thickness) as initial templates, followed by Au seed attachment (~2.0 nm in diameter). During polymer coating, the SiO2 shells were etched away, and defined gaps were formed between the Fe3O4 nanorods and polymer shells. Seeded growth was performed inside the gaps to prepare the hybrid nanorods, which comprise one Au nanorod and one Fe3O4 nanorod in a parallel configuration, as shown in the transmission electron microscopy (TEM) images (Fig. 2A) and elemental mapping (fig. S19). Because of the confinement of the poly- mer shells, radial growth was limited, and preferable longitudinal growth produced the Au nanorods. This growth mode allows easy tuning of the nanorod length (fig. S18, B to E) and shifts the LSPR of the Au nanorods from 560 to 880 nm (fig. S18F). The parallel align- ment of the hybrid nanorods allowed the Au nanorods to assemble into chiral superstruc- tures under a chiral magnetic field and pro- duce CD responses (fig. S20, A and B). Switching the magnet dipole did not alter the CD profile, but changing the magnet position to the opposite side of the sample produced a CD spectrum with an opposite sign (fig. S20C). Applying two identical magnets in their at- traction configuration generated a parallel magnetic field with a reduced field gradient between the two magnets and reduced the CD signals (fig. S20, D and E). The disappearance of CD signals confirmed that linear super- structures assembled in a uniform magnetic field could not produce CD signals in our ex- perimental conditions. We also fixed the hybrid nanorods under the absence and presence of uniform and chiral magnetic fields using a photocuring polymerization method, which produced random, linear, and chiral struc- tures, respectively. Although films containing random and linear structures were not op- tically active, the film with chiral superstructures had CD signals. None of the three structures showed evident responses to a magnet once fixed in polymer films (fig. S21). These exper- iments suggested that the observed CD re- sponses in a single permanent magnet were not induced by extrinsic chirality, a CD pro- perty of achiral superstructures (44), or by mag- netic circular dichroism (45). We measured the CD spectra of Au nano- rods while changing the position of the mag- net vertically (Fig. 2B and fig. S22, A and B). The CD signal declined gradually after the magnet was moved from −1.25 to 0 cm along the z axis (vertical direction) (fig. S22, C and D), and the spectra changed the sign across 0 cm (Fig. 2C) in response to the field chirality changes shown in fig. S22, E and F. The dependence of CD signals on magnet posi- tion was similar when the magnet dipole was along the z axis (fig. S23). Changing the samples’ position relative to the magnet was equivalent but induced a decrease in CD intensity when the sample-magnet separation increased up to 3.0 cm along the x axis (fig. S24). The CD spectra of Au nanorods were then measured by decreasing field strength from 31.2 to 25.1, 18.3, 12.9, and 5.5 mT, which corresponded to the magnet being 2.3, 2.5, 3, 4, and 5 cm away from the nanorod dispersion. We observed a consistent decrease in CD intensity for Au nanorods with different LSPR positions and for left- and right-handed chiral superstructures (Fig. 2, D and E, and fig. S25), which cor- responded to the decrease in field rotation when the magnet departed the sample (Fig. 2F). Detailed field analysis (Fig. 2, G to J) indicated that the Dw decreased consistently in the four chiral domains for increasing magnet-sample separation. The sample in the first quadrant was in a left-handed field in the By field (Fig. 2, G and H), which caused the negative signals in CD spectra. In the Bz field, the chiral field in the first quadrant produced positive CD signals (Fig. 2, I and J). The peak intensities of ex- tinction and CD of the hybrid nanorods were linearly proportional to Au concentration at fixed magnet and sample positions (fig. S26), which was consistent with Beer’s law and molar ellipticity in CD. To verify the chirop- tical properties, we modeled the chiral super- structures and calculated their CD spectra using a finite element method. The simulated spectra in fig. S27 demonstrated CD responses of Au nanorods with large separations. Considering the lack of nanorod positional order in exper- iments, complex models were further developed to resolve the structures in figs. S15 and S17. These models contain nanorods with random Li et al., Science 380, 1384–1390 (2023) 30 June 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E A B Fe3O4/Au z (0,2.5,1.25) (0,2.5,0) 50 nm Polymer (0,2.5,-1.25) 1.25 cm -1.25 cm C 400 200 0 -200 ) g e d m ( D C y k x D ) g e d m ( D C 400 300 200 100 0 -100 -200 -300 -400 31.2 mT 25.1 mT 18.3 mT 12.9 mT 5.5 mT By 5.5 mT Bz 12.9 mT 18.3 mT 25.1 mT 31.2 mT 200 300 400 500 600 700 800 900 Wavelength (nm) G ) g e d ( 3 0 H −3 0 −2 ) m c ( Z E n e y p B r e P l a c z i t B r e V 7 6 5 4 3 2 1 1 2 3 4 5 6 7 -455.0 F ) g e d m ( D C 5.5 12.9 18.3 25.1 31.2 Field Strength (mT) 455.0 I ) g e d ( 3 0 −3 0 −2 J ) m c ( Z -400 200 300 400 500 600 700 800 900 Wavelength (nm) 2.3 cm 3 cm 5 cm 6.1 −4 −6 −4 −2 0 Y (cm) 2 4 6 −4 −6 −4 −2 2 4 −6.1 6 0 Y (cm) Fig. 2. Magnetic assembly of Fe3O4/Au hybrid nanorods into chiral super- structures. (A) TEM image of the Fe3O4/Au hybrid nanorods wrapped within polymer shells. (Inset) Schematic structure of the Fe3O4/Au hybrid nanorods. (B) Schematic illustration of the magnet position during the CD measurement. (C) CD spectra of the hybrid nanorods measured by changing the magnet position. (D) CD spectra of the hybrid nanorods under magnetic fields with consistent direction and decreasing strength. (E) Dependence of CD intensity on the field strength and the rod aspect ratios. The magnetic fields are defined as By and Bz when the magnet dipole is parallel to y and z axis, respectively. (F) Simulated field distributions of the cubic magnet at given distances to the magnet surface. (G and I) Rotation angles of magnetic fields between two yz cross sections, (–0.2, y, z) and (–0.8, y, z). The orientation of the magnet is shown in the inset in each plot. (H and J) The corresponding field rotation angles (color map) and the local magnetic field at the cross section (–0.2, y, z). The magnet has a horizontal magnetic dipole in (G) and (H) and a vertical magnetic dipole in (I) and (J). The field rotation is calculated by Dw(y, z) = w(–0.8, y, z) − w(–0.2, y, z). positions but constant rotation in different planes (fig. S28), resembling the nanorod align- ment in different cross sections in the SEM im- ages. The simulated CD spectra show rotation angle–dependent CD responses, which explains the decrease of CD intensity with the rotation angles of magnetic fields. We also observed similar dependence in nanorods of increas- ing aspect ratios, and the continuous redshift of CD peaks in fig. S29 is qualitatively consistent with the measured CD spectra in fig. S25. Changing the directions of the magnetic field produced more complex CD responses. We considered the rotation of the magnet within the xy and yz planes to explain the involved mechanisms (Fig. 3A). The alignment of the magnetic dipole of the magnet relative to the axes was defined by the azimuth angle q. The CD spectra showed intensity changes when q increased to 180° in the xy plane (Fig. 3B), which increased to a saturated value and decreased again (fig. S30). Given that the magnet was rotated in the xy plane (fig. S31), the local magnetic fields and the field distri- bution were analyzed for different q values. The schemes in fig. S31 illustrate the magnet’s constant position and varying orientation. Because the incident light was along the x axis, the magnetic fields within the mea- sured domains in the yz and xz planes were further plotted in Fig. 3, C and D, respectively. We observed only slight changes in field di- rection and field distribution within the yz plane for different magnet orientations (Fig. 3C). In the xy planes, however, the field dis- tribution exhibited dramatic changes when q increased to 90°, which led to excitation of longitudinal mode with gradually increased strength (Fig. 3D). For incident light along the x axis, the plasmonic excitation of Au nano- rods was determined by and could be predicted from the nanorod alignment, light incidence, and light polarization (46). Because of the changes in magnetic field direction and dis- tribution shown in fig. S32, there was an associated change in the LSPR excitation of the Au nanorods when q increased to 90°. At 0°, the field was nearly parallel to the x axis, and the resulting parallel alignment of nanorods to the incident light suppressed the longitu- dinal mode. At 90°, the magnetic field being nearly parallel to the z axis led to the exci- tation of the longitudinal mode. We could predict the longitudinal extinction of the Au nanorods through the equation sin2(ϕ), where ϕ is the angle between the local-field direction and the light incident direction. Figure 3E shows a symmetric trend with maxi- mum extinction between 60 and 120°. Compar- ison of the predicted longitudinal extinction Li et al., Science 380, 1384–1390 (2023) 30 June 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E A C D y y 0o x z in x-y y in y-z x k (0,-2,-0.5) z z 0o y y 48.00 B 180 150 120 90 60 30 ) g e d ( l e g n A 0 200 300 400 500 600 700 800 900 Wavelength (nm) -147.0 0.5 0.0 ) m c ( Z -0.5 -1.0 0.5 ) m c ( Z 0.0 -0.5 -0.5 -0.5 Y (cm) 0.0 -1.0 -0.5 Y (cm) 0.0 -1.0 -0.5 Y (cm) 0.0 -1.0 -0.5 Y (cm) E -35 -90 0.0 F 1 0 ) g e d m ( D C 50 0 -50 -100 -150 0.0 X (cm) 0.5 -0.5 0.0 X (cm) 0.5 -0.5 0.0 X (cm) 0.5 -0.5 0.0 X (cm) 0.5 690 nm 535 nm 0 30 60 90 120 150 180 Angle (deg) Fig. 3. Tunable CD spectra of the assembled superstructures. (A) Schematic illustration of the azimuth changes of a cubic permanent magnet within the xy and yz plane during CD measurements. (B) CD spectra of the hybrid nanorods measured by changing the magnet azimuth in the xy plane. (C and D) The magnetic field of the cubic magnet in (C) yz and (D) xz planes. The azimuth of the magnet is 0° to 30°, 60°, and 90° from left to right. (E) Predicted longitudinal extinction of the nanorods based on the analytical solution. (F) Dependence of the CD intensity on magnetic field q. with the CD intensity measured at different q (Fig. 3F) suggested that CD intensity would depend on the magnet q and that the CD responses of the assembled superstructures would be determined by the longitudinal ex- tinction changes when the magnet azimuth increased within the xy plane. Changing the magnet azimuth in the yz plane altered the field chirality and led to a different CD response. The measured CD spec- tra are shown in fig. S33. The CD peaks at 545 and 698 nm simultaneously switch their signs at about 20° and 110° during the measurement (Fig. 4A, color map), which suggested changes in handedness of the assembled superstruc- tures. To verify this hypothesis, we used the analytical model to directly map the local-field rotation direction, chirality, and handedness (fig. S2). At q = 0°, the sample was 2 cm away from the magnet center with an upward off- set of 0.5 cm, which led to an azimuth of ~14° relative to the magnet center (fig. S34). Understanding the CD spectrum at q = 0° re- quired access to field properties at this spe- cific sample location in the yz plane. The CD changes in response to magnet ro- tation were studied by analyzing the local- field properties in different q. In Fig. 4B, the plots of field distribution and field chirality within the yz plane showed a similar quadru- pole field chirality. The normalized field vectors at x = –0.2 and –0.8 cm were superimposed in Fig. 4B to illustrate local-field rotation. The field rotation angles along a trajectory high- lighted in fig. S34C are shown in Fig. 4C, which correspond to the field changes during experimental measurement. The Dw was ini- tially positive but changed its sign at q = 15° and 105°, which is consistent with the experi- mental chirality-transition angles plotted in Fig. 4D. The Pearson correlation coefficients between the field rotation angles and CD intensity at 698 and 545 nm are –0.989 and 0.987, respectively (fig. S35). The strong negative and positive correlation indicates that the field chirality changes could explain the dependence of the superstructure handedness and CD spec- tra on magnet rotation in the yz plane. Optical rotatory dispersion We also studied the optical rotatory disper- sion (ORD), which measures the polarization rotation of a linearly polarized light. Left- and right-handed circularly polarized light interacts differently with chiral structures and travels at different speeds inside of them. Because linearly polarized light comprises two circu- larly polarized light beams with the same magnitude but opposite handedness, these two highly correlated beams will develop a phase difference, leading to the polarization rotation of the incident beam (Fig. 4E). The ORD effect was tested by applying an analyzer to the incident beam and observing the color changes. Only light of a specific wavelength can pass through the analyzer at a polarization angle (a) when the material is optically active. Experimentally, a is defined as the angle between the analyzer polarization direction and the horizontal baseline, and the polariza- tion of the polarizer is fixed along the vertical direction. Digital images of a nanorod dispersion are shown in Fig. 4F, before (left) and after (right) application of a z-directional magnetic field at a = 3°. The original dispersion appeared dark without noticeable colors because only mini- mal light could transmit through the analyzer. Under a z-directional magnetic field, the top domain turned yellow, and the bottom turned red, demonstrating the formation of chiral superstructures with opposite handedness in these two domains, which is consistent with the predicted field chirality in the first Li et al., Science 380, 1384–1390 (2023) 30 June 2023 4 of 7 1 0 -1 -2 -3 -4 -5 -3 -2 -1 0 Y (cm) 1 2 3 C 6.1 D ) g e d m ( D C 200 100 0 -100 -200 -300 −6.1 698 nm 545 nm 0 30 60 90 120 150 180 Angle (deg) -30o -27o -24o -21o -18o -15o -12o -9o -6o -3o 0o 3o 6o 9o 12o 15o 18o 21o 24o 27o 30o RES EARCH | R E S E A R C H A R T I C L E A ) g e d ( l e g n A 180 150 120 90 60 30 B 180.0 ) m c ( Z 0 200 300 400 500 600 700 800 900 Wavelength (nm) -255.0 E Polarizer z G Analyzer α x F y (Bz) w/o mag w/ mag (Bz) By Bz Bx No Fig. 4. Tunable CD spectra and optical rotatory dispersion of the assembled superstructures. (A) CD spectra of the hybrid nanorods measured by changing the magnet azimuth in the yz plane. (B) The magnetic field rotation (color map) and the normalized field vectors (arrows) of the cubic magnet. The magnetic field rotation is calculated by Dw(y, z) = w(–0.8, y, z) – w(–0.2, y, z). The normalized magnetic field vectors in (–0.8, y, z) are indicated with black arrows, and the field vectors in (–0.2, y, z) are indicated with orange arrows. (C) Dependence of field rotation on the magnet field azimuth. (D) Dependence of the CD peak intensity on the magnet field azimuth. (E) Schematic illustration of the ORD measurement. The a is introduced as the angle between the polarization direction of the analyzer and the horizontal direction. (F) Digital images of a hybrid nanorod dispersion in a cuvette without (w/o mag) and with (w/ mag) a magnetic field. The a is 3°. (G) Digital pictures of the hybrid nanorod dispersion under different magnetic field conditions. The analyzer’s a was switched from –30 to 30° during the measurement. and fourth quadrants of the magnet. This ORD effect was determined by superstruc- ture handedness and the angle a. Under a z-directional magnetic field, we observed sim- ilar two-color domains when a increased from –30 to 30° (Fig. 4G), with red-orange-yellow- green changes in the top domain and oppo- site color changes in the bottom domain. Changing the magnetic field to the y axis led to color switching between the two domains, corresponding to a field chirality transition. Nanorod dispersion under the absence and presence of the x-directional magnetic field only exhibited contrast changes at different a because of the negligible CD responses at these two conditions. Generalizing the all-scale chiral assembly The chirality formed by nanoscale magnetic assembly could be transferred to organic molecules, polymers, oxides, and semiconduc- tors. These guest materials were introduced to the magnetic nanorods through surface-coating and doping methods, which have the advan- tages of wide material accessibility and easy further processing. Starting with Fe3O4@SiO2 nanorods with a length of 107.6 ± 5.2 nm, a diameter of 13.0 ± 1.7 nm, and silica thickness of 5.0 ± 0.5 nm, we coated their surface with Cu2O nanoparticles and resorcinol-formaldehyde (RF), with the latter being converted into MnO2 by reacting with KMnO4 (47–49). The resulting Fe3O4@SiO2@MnO2 nanorods (Fig. 5A) under- went oxidation-induced volume expansion and created nanogaps between porous MnO2 nano- shells and Fe3O4@SiO2 cores. The extinction spectra of these samples in Fig. 5B showed broad extinction of Fe3O4@SiO2 and Fe3O4@ SiO2@Cu2O nanorods, and extinction peaks of Fe3O4@SiO2@RF and Fe3O4@SiO2@MnO2 nanorods. These four samples showed evident CD activities under magnetic fields that can also be tuned by changing the field directions (Fig. 5C). The CD peak positions being near their extinction peak positions indicates that the chirality in the superstructures transferred from the host nanorods to the guest mole- cules. In addition, changing the field direction from the y axis to the z axis changed the chiral superstructures from left-handed to right-handed symmetry. Transfer of chirality to small molecules was demonstrated by doping organic dyes into RF polymeric shells through electrostatic inter- actions. Figure S36A shows a typical TEM image of the Fe3O4@SiO2@RF nanorods with 100-nm RF shells that carry negative charges after a condensation reaction. We chose three dyes—methylene blue, methylene green, and neutral red—and their molecular structures are given in fig. S36B. These dyes develop positive charges after dissociation of chloride anions in water and can be doped into porous RF shells by mixing with the nanorods at room temperature. The dyed solutions appeared blue, green, and red after removing the excess molecules, with extinction spectra exhibiting distinct peaks after successful doping (Fig. 5D). Using the green dye–doped nanorods as an example, we confirmed their CD responses under different field directions (fig. S36C), with negligible CD signals under no and x-directional magnets and opposite CD peaks in y- and z-directional magnetic fields. The CD spectra of nanorods doped with all three dyes (Fig. 5E) showed peaks at their charac- teristic wavelengths. The observation is con- sistent with that of plasmonic nanorods and demonstrated the formation of chiral super- structures and the successful transfer of chi- rality from the nanoscale to molecular levels. The g factor calculated in Fig. 5F has a max- imum of ~0.003 for methylene blue–doped nanorods, a value comparable to that of classic chiral molecules and magnetically assembled chiral superstructures of inorganic molecules, polymers, and semiconductors (figs. S37 and Li et al., Science 380, 1384–1390 (2023) 30 June 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E B E 200 nm Fe3O4@RF Methylene blue Methylene green Neutral red A D 2.5 2.0 1.5 1.0 0.5 . ) . u a ( e c n a b r o s b A 0.0 300 400 500 600 700 800 900 1000 400 300 200 100 0 -100 -200 -300 ) g e d m ( D C -400 ) . u . a ( e c n a b r o s b a d e z i l a m r o N Fe3O4@SiO2 Fe3O4@SiO2@CuO2 Fe3O4@SiO2@RF Fe3O4@SiO2@MnO2 1.0 0.8 0.6 0.4 0.2 C ) g e d m ( D C d e z i l a m r o N 0.0 300 400 500 600 700 800 900 Wavelength (nm) F 1 0 -1 Fe3O4@SiO2/By Fe3O4@SiO2/Bz Fe3O4@SiO2@Cu2O/By Fe3O4@SiO2@Cu2O/Bz Fe3O4@SiO2@RF/By Fe3O4@SiO2@RF/Bz Fe3O4@SiO2@MnO2/By Fe3O4@SiO2@MnO2/Bz 300 400 500 600 700 800 900 Wavelength (nm) undoped/By undoped/Bz blue/By blue/Bz green/By green/Bz red/By red/Bz Methylene blue Methylene green Neutral red 7 6 5 4 3 2 1 ) 3 - 0 1 ( r o t c a f - g 0 300 400 500 600 700 800 900 300 400 500 600 700 800 900 Wavelength (nm) Wavelength (nm) Wavelength (nm) Fig. 5. All-scale magnetic assembly of achiral molecules into chiral superstructures. (A) TEM image of the Fe3O4@SiO2@MnO2 core-shell nanorods. (B) Extinction spectra of core-shell Fe3O4@SiO2 and Fe3O4@SiO2@RF nanorods, Fe3O4@MnO2 yolk-shell nanorods, and Fe3O4@SiO2@Cu2O core-satellite nanorods. (C) The corresponding CD spectra measured under y- and z-directional magnetic fields. (D) Extinction spectra of the Fe3O4@SiO2@RF nanorods and nanorods doped with the three organic dyes. (Insets, left to right) Digital pictures of the nanorods before and after doping with methylene blue, methylene green, and neutral red. (E) CD spectra of the doped and undoped nanorods under y- and z-directional magnetic fields. (F) The g factor of the doped nanorods under a z-directional magnetic field. S38). We further demonstrated that the mag- netic assembly strategy could be extended to the assembly of fluorophores for generat- ing circularly polarized luminescence. When europium-doped NaYF4 nanorods were decor- ated with Fe3O4 nanoparticles, the fluorescent nanorods could be assembled into chiral superstructures and exhibited circularly polar- ized luminescence (fig. S39). Discussion The consistent rotation of the local-field vec- tors of the cubic permanent magnet forms the quadrupole field chirality with alternat- ing left-handed and right-handed magnetic fields in the four quadrants. Such chiral mag- netic fields induce the assembly of magnetic nanorods into chiral superstructures, with handedness and chirality determined by the local features of the magnetic fields. This strategy allows remote, reversible, and instan- taneous assembly of chiral superstructures from nanostructures of various chemical com- pounds (plasmonic materials, polymers, oxides, metals, semiconductors, fluorescent nanostruc- tures, and molecular moieties) and active tuning of their CD responses in a broad range of spectra and circularly polarized lumines- cence, as long as they can be properly bound to the magnetic nanorods. Fixing the chirality of these chemical compounds at all scales is possible by embedding the formed super- structures in polymer substrates, which could be realized by applying an external magnetic field during polymerization. This simple strategy makes the chiral superstructures nonvolatile without external magnetic fields and highly accessible for portable chiroptical devices. RE FERENCES AND NOTES 1. G. Singh et al., Science 345, 1149–1153 (2014). 2. X. Lan et al., J. Am. Chem. Soc. 137, 457–462 (2015). 3. W. Ma et al., Chem. Rev. 117, 8041–8093 (2017). J. Lv et al., Nat. Rev. Chem. 6, 125–145 (2022). 4. 5. M. Hentschel, M. Schäferling, X. Duan, H. Giessen, N. Liu, Sci. Adv. 3, e1602735 (2017). 6. C. Zhou, X. Duan, N. Liu, Acc. Chem. Res. 50, 2906–2914 (2017). 7. R. Schreiber et al., Nat. Commun. 4, 2948 (2013). 8. D. Zerrouki, J. Baudry, D. Pine, P. Chaikin, J. Bibette, Nature 455, 380–382 (2008). 16. C. Hao, L. Xu, H. Kuang, C. Xu, Adv. Mater. 32, e1802075 (2020). 17. F. Neubrech, M. Hentschel, N. Liu, Adv. Mater. 32, e1905640 (2020). 18. Y. Y. Lee, R. M. Kim, S. W. Im, M. Balamurugan, K. T. Nam, Nanoscale 12, 58–66 (2020). 19. Z. Li, Q. Fan, Y. Yin, Chem. Rev. 122, 4976–5067 (2022). 20. N. J. Greenfield, Nat. Protoc. 1, 2876–2890 (2006). 21. B. Shen et al., Sci. Adv. 4, eaap8978 (2018). 22. B. Frank et al., ACS Nano 7, 6321–6329 (2013). 23. A. Kuzyk et al., Nature 483, 311–314 (2012). 24. D. Winogradoff et al., Adv. Sci. (Weinh.) 8, 2003113 (2021). 25. K. Martens et al., Nat. Commun. 12, 2025 (2021). 26. Y. Chen et al., Nat. Rev. Phys. 4, 113–124 (2022). 27. P. Zhan et al., Sci. Adv. 5, eaax6023 (2019). 28. L. Xin, C. Zhou, X. Duan, N. Liu, Nat. Commun. 10, 5394 (2019). 29. S. Li et al., J. Am. Chem. Soc. 138, 306–312 (2016). 30. A. Cecconello, L. V. Besteiro, A. O. Govorov, I. Willner, Nat. Rev. Mater. 2, 1–19 (2017). 31. A. Kuzyk, M. J. Urban, A. Idili, F. Ricci, N. Liu, Sci. Adv. 3, e1602803 (2017). 32. S. Zhou et al., Nature 612, 259–265 (2022). 33. Z. Cheng, M. R. Jones, Nat. Commun. 13, 4207 (2022). 34. J. Lu et al., Science 371, 1368–1374 (2021). 35. G. M. Whitesides, B. Grzybowski, Science 295, 2418–2421 (2002). 9. L. Xu et al., Nature 601, 366–373 (2022). 10. N. A. Kotov, L. M. Liz-Marzan, P. S. Weiss, ACS Nano 15, 36. X. Lan, Q. Wang, Adv. Mater. 28, 10499–10507 (2016). 37. C. A. Batista, R. G. Larson, N. A. Kotov, Science 350, 1242477 12457–12460 (2021). 11. K.-J. Jeong et al., ACS Nano 14, 7152–7160 (2020). 12. H.-E. Lee et al., Nature 556, 360–365 (2018). 13. W. Jiang et al., Science 368, 642–648 (2020). 14. G. González-Rubio et al., Science 368, 1472–1477 (2020). 15. L. Ohnoutek et al., Nat. Photonics 16, 126–133 (2022). (2015). 38. B. N. Khlebtsov, V. A. Khanadeev, N. G. Khlebtsov, Langmuir 24, 8964–8970 (2008). 39. W. Xu et al., Nano Lett. 17, 2713–2718 (2017). 40. Z. Li, C. Qian, W. Xu, C. Zhu, Y. Yin, Sci. Adv. 7, eabh1289 (2021). Li et al., Science 380, 1384–1390 (2023) 30 June 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E 41. M. Wang, Y. Yin, J. Am. Chem. Soc. 138, 6315–6323 ACKN OWLED GMEN TS (2016). 42. Z. Li et al., J. Am. Chem. Soc. 143, 15791–15799 (2021). 43. Z. Li, J. Jin, F. Yang, N. Song, Y. Yin, Nat. Commun. 11, 2883 (2020). 44. E. Plum, V. Fedotov, N. Zheludev, Appl. Phys. Lett. 93, 191911 (2008). 45. B. Han, X. Gao, J. Lv, Z. Tang, Adv. Mater. 32, e1801491 (2020). 46. M. Wang et al., J. Am. Chem. Soc. 135, 15302–15305 (2013). 47. C. Liu, H. Dong, N. Wu, Y. Cao, X. Zhang, ACS Appl. Mater. Interfaces 10, 6991–7002 (2018). 48. S. Zhou et al., Nanoscale 12, 15460–15465 (2020). 49. Y. Bai, X. Yao, X. Wang, Y. Yin, ChemNanoMat 6, 1186–1190 (2020). 50. Z. Li, Programming code for: A magnetic assembly approach to chiral superstructures, version 1, Zenodo (2023); https://doi. org/10.5281/zenodo.7686903 We thank the Central Facility for Advanced Microscopy and Microanalysis at University of California Riverside (UCR) for help with TEM analysis and G. Ung’s group at University of Connecticut for assistance with measuring the circularly polarized luminescence. Funding: US National Science Foundation CHE- 2203972 (Y.Y.). Authors contributions: Z.L. and Y.Y. conceived and planned this study. Z.L. and Q.F. developed the methodology for modeling and simulation. Z.L. developed the synthesis and characterization of nanoparticles. Z.L., Z.Y., and C.W. performed the CD measurements. Z.W. contributed to the circularly polarized luminescence studies. All authors discussed and analyzed the results and contributed to the writing of this paper. Competing interests: Y.Y. and Z.L. have filed a patent application related to this work filed by UCR. The anthors declare that they have no other completing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper or the supplementary materials. The programming code to model magnetic fields and datasets used to analyze the field chirality are available online at Zenodo (50). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg2657 Materials and Methods Supplementary Text Figs. S1 to S39 Table S1 References (51–57) Submitted 12 December 2022; accepted 22 May 2023 10.1126/science.adg2657 Li et al., Science 380, 1384–1390 (2023) 30 June 2023 7 of 7
10.1126_science.adg0059
RES EARCH IONTRONICS Cascade-heterogated biphasic gel iontronics for electronic-to-multi-ionic signal transmission Weipeng Chen1,2†, Linxin Zhai3†, Suli Zhang4,5†, Ziguang Zhao1,2†*, Yuhao Hu1,2, Yun Xiang1, Huirong Liu4,5, Zhiping Xu3, Lei Jiang1,2, Liping Wen1,2* Currently, electronics and iontronics in abiotic-biotic systems can only use electrons and single-species ions as unitary signal carriers. Thus, a mechanism of gating transmission for multiple biosignals in such devices is needed to match and modulate complex aqueous-phase biological systems. Here we report the use of cascade-heterogated biphasic gel iontronics to achieve diverse electronic-to-multi- ionic signal transmission. The cascade-heterogated property determined the transfer free energy barriers experienced by ions and ionic hydration-dehydration states under an electric potential field, fundamentally enhancing the distinction of cross-interface transmission between different ions by several orders of magnitude. Such heterogated or chemical-heterogated iontronics with programmable features can be coupled with multi-ion cross-interface mobilities for hierarchical and selective cross-stage signal transmission. We expect that such iontronics would be ideal candidates for a variety of biotechnology applications. I n biological systems, neural networks with complex morphologies and highly polar- ized interfacial architectures can support elaborate bioionic and biochemical signal communications between different neu- rons (1, 2). Such structures and functions of complex neural networks provide inspiration for designing cascade-gated architectures ca- pable of implementing neuron-like multivariate ionic or chemical signal transport that could interface with and modulate physiological processes. Recently, artificial electronic and iontronic systems, which broke the information barriers between abiotic and biotic interfaces, have at- tracted considerable attention for application in biosensors (3–6), neuroprosthetics (7, 8), and implantable neuromorphic devices (9–12). Electronics, especially neuromorphic electron- ics (11–14), directly use programmable electrical impulse signals to treat biological interfaces as signal transducers for modulating physiologi- cal activities (14–16), whereas iontronics ex- hibit the capability of individually performing the conversion between intrinsic ionic and elec- trical signals for abiotic-biotic systems (17, 18). However, compared with biological neuronal networks with diverse biosignal transmission invoked by neural action potentials (fig. S1), 1CAS Key Laboratory of Bio-inspired Materials and Interfacial Science, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, P. R. China. 2School of Future Technology, University of Chinese Academy of Sciences, Beijing 100049, P. R. China. 3Applied Mechanics Laboratory, Department of Engineering Mechanics, Tsinghua University, Beijing 100084, P. R. China. 4Department of Physiology and Pathophysiology, School of Basic Medical Sciences, Capital Medical University, Beijing 100069, P. R. China. 5Beijing Key Laboratory of Metabolic Disorder Related Cardiovascular Disease, Capital Medical University, Beijing 100069, P. R. China. *Corresponding author. Email: zhaoziguang@ucas.ac.cn (Z.Z.); wen@mail.ipc.ac.cn (L.W.) †These authors contributed equally to this work. most electronics and iontronics are still re- stricted by the single-species attributes of signal carriers (electrons or single ions) that cannot carry more biocompatible information. These existing iontronics constructed from conven- tional gated or nongated materials do not faith- fully mimic multi-ionic cotransport, thereby limiting their characteristic signal expression in matching biological tissues (19). Ionic cur- rent signal transmission does not even discrim- inate between various ions in hydrogel-based iontronics (20–22). In the realm of iontronics, diverse electronic-to-multi-ionic transmission and processing in aqueous-phase biological environments remain a challenge. Design and fabrication of cascade-heterogated biphasic gels Here we report the use of a biphasic gel iontronic with cascaded heterointerface–gated properties to construct diverse ionic cross-medium trans- mission as a signal communication principle (Fig. 1A). The cascade-heterogated biphasic gel (HBG) has phase-separated heteronetwork structures that incorporate opposite binary gel phases: a microscale ion-enriched internal gel phase (IE phase) and a low-conductivity contin- uous gel phase (LC phase) (Fig. 1B). Specifically, the IE phase features hydrogel microinclu- sions that can store and transmit hydrated ions, and the LC phase serves as an organogel matrix that functions as the partially dehy- drated ion transmission medium. Ionic cross- medium transmission can be triggered by external stimulation of both continuous- and pulse-based electrical signals, during which partial shedding and reconfiguration of cat- ionic hydration shells, acting as ionic partial hydration and rehydration states (Fig. 1C), must occur alternately and consecutively. Thus, biphasic multi-interfaces induce cascade- heterogated effects that determine how ions experience transfer free energy barriers (DE) and local driving force (º DV) under an elec- tric potential (Fig. 1, D and E). Because of the association of these factors with ionic hydration- dehydration energies, the distinction of cross- interface transmission between multivariate cations (X+ and Xn+) is enhanced by several orders of magnitude. By way of the sorting and control of ionic transfer energy barriers at heterogated interfaces, the HBG system can process electronic-to-multi-ionic hierar- chical transmission and selective ionic cross- stage transmission. An in situ phase-separated polymerization strategy was introduced to fabricate HBG iontronics in which interfacial covalent inter- actions could maintain stable heterogated ar- chitectures (fig. S2). Heterogeneous structures of HBGs with different phase compositions were confirmed by confocal laser scanning microscopy (CLSM) (Fig. 1B and fig. S3) and scanning electron microscopy (SEM) (fig. S4). In CLSM images, it was observed that IE micro- inclusions (blue-stained) were uniformly dis- persed in the LC phase (orange-stained) and found that their average diameters increased from 1.48 to 2.67 mm upon increasing the ratio of the IE phase (fig. S5). These results indicate that cascade-heterogated structures are modu- lated by the microphase separation degree. Ion species and concentrations did not affect HBG heterostructure (figs. S6 and S7). These samples were denoted as HBG-y, where y rep- resents the volume ratio of the LC phase to the IE phase (fixed to 1). Desirable mechan- ical properties and environmental stability of HBG iontronics ensure high-reliability ion trans- port in multimedia-based environments (figs. S8 and S9). Ionic transmission characteristics of cascade-heterogated biphasic gels In nongated hydrogel iontronics, X+ and Xn+ cations are highly hydrated and can contin- uously move with extremely low viscous restric- tions in water-rich environments and easily traverse the neutral hydrogel network in an electric potential field (Fig. 2A). There is no obvious differentiation between the DE of hy- drated X+ and Xn+ cations during homogeneous transmission (fig. S10). In HBG systems, the cascade-heterogated mechanism can effective- ly govern ion transport owing to the ionic in- trinsic DE under the external stimulation of continuous or pulse-based electrical signals (Fig. 2B). In the cascaded cross-interface ion transmission, the partial shedding and recon- figuration of cation hydration shells, which correspond to the ionic partial hydration and rehydration states, occurs alternately and suc- cessively. To further illustrate the heterogated effect derived from asymmetric chemical and spatial structures, we established the molec- ular dynamics model of the heterointerface Chen et al., Science 382, 559–565 (2023) 3 November 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E A D B Ion-enriched internal phase in int IE phase Lo L Low-conductive ow con nt continuous phase ( (LC phase) CLSM image CLSM image C Heterogated transmission Electrical signal Electrical signalg IE phase Hydrated ion2 d ie Xn+ (Xn+) Hydrated ion1 e ed Hydrateed ion ) (X+) Bi h i Biphasic-gel Iontronics Biphasic-gel Iontronics l I t i e r u t c u r t S n o i t u b i r t s d i Aqueous phase Oleophilic network Hydrophilic network Oleophilic phase IE phase LC phase E CCCr CroCross-ss ii aaa ans imis ision inii tntn erfferfaceace trtra ssmm oonnn nn ffrfaff Cascade-heterogated network structure PP Partially P hydrated ion2 dr (Xn+) Partially h on hydrated ion1 (X+) LC phase Ionic signal Transfer energy barrier ( E) Local driving force ( V) Hydration Partial hydration Rehydration Hydration Partial hydration Rehydration Hydrated ion Partially hydrated ion EXn+ EX+ y g r e n e e e r F r IE >> r Hydrated ion > r LC > r Partially hydrated ion Analog ionic transmission path through the heterogated interface IE phase 0 LC phase IE phase Coordinate e g a t l o V VX+ VXn+ IE phase LC phase 0 Coordinate IE phase Fig. 1. Heterogeneous structures and cross-interface ion transmission of cascade-heterogated biphasic gel iontronics. (A) Structural illustration and (B) CLSM image of the HBGs with phase-separation heteronetworks. Scale bar, 20 mm. (C) Heterogated cross-interface ion transmission. Owing to the distinct hydration-dehydration energies between different X+ and Xn+, the cascade-heterogated effects fundamentally enlarged the hierarchy discrepancy of the ionic transfer energy barriers to construct diverse electronic-to-multi-ionic signal transmission by the application of electrical signals. (D) Analysis of the network structure distribution at the heterogated interface and the analog ionic transmission path through the heterogated interface from the hydrated state in the IE phase to the partially hydrated state in the LC phase. (E) The obvious distinction of transfer energy barrier (DE) and the local driving forces (º DV) of X+ and Xn+ featuring the hydration and partial hydration states across the heterogeneous interfaces. to simulate the ionic transmission path and the related transfer free energy barrier (fig. S11). Crossing the heterointerface from the IE phase to the LC phase, the attractive and repulsive interactions between cation hydra- tion shells and IE and LC phases, as well as the steric limitations of heterogated interfaces, led to a substantial increase in DE. Meanwhile, the resultant transport resistance could be counteracted by the local driving force from the electric field, which was reflected by the notable DV. Ionic transmission paths in the IE phase with an average estimated size of 6.60 to 7.26 nm were found to be larger than those of hydrated ions, and so a partial de- hydration transformation reduced the effec- tive size of hydrated ions, allowing them to access the LC phase, in which the estimated path size of 1.5 to 3.0 Å was larger than that of the partially hydrated ions (Fig. 1D and fig. S12A). The heterogated interface with a high degree of permeation also acted as a transi- tion layer to reduce the driving force of ions across the heterointerface, and the interfa- cial gating thickness did not affect the ionic cross-interface energy consumption originating from its hydration-dehydration transforma- tion (fig. S13). Subsequently, an energy com- pensation existed for the disengagement from the attractive interaction between the ion and the IE phase, and the partially hydrated ion with a related high transfer energy in the LC phase was relayed to the next IE microinclu- sions, accompanied by an energetically favor- able reconfiguration to realize ion rehydration. For one unit of cascaded cross-interface ion transport (fig. S12B), the ratios of K+, Ca2+, and Fe3+ hydration-dehydration energies and ion mobilities between the biphasic system were simulated to evaluate different cross-interface Þ of the DE values DE three typical ions (fig. S12, C and D). Specifi- cally, ionic radius, electrostatic force, hydrogen bonding, and other weak interactions, which Ca2þ > DEKþ Fe3þ > DE ð together affect the stability of the ionic hydra- tion shell, were believed to be intrinsic factors in determining the ionic hydration-dehydration energy (23, 24). Theoretically, ions with high charge densities can lead to the strong electro- static ordering of nearby water molecules to stabilize their hydration shells, thus necessi- tating the higher hydration-dehydration ener- gies to overcome cross-interface transfer energy barriers (25). Figure 2, C and D, and fig. S14 show similar ionic current variations of nongated hydrogels containing 1 mM K+, Ca2+, or Fe3+ under dif- ferent voltage gradients, and their comparable transfer potentials were confirmed in terms of the corresponding voltage thresholds (all 1 mV). These indiscriminate results demon- strate that homogeneous networks of non- gated hydrogels had difficulty recognizing ionic intrinsic features for the distinction of ion transport. In contrast, the ionic current prop- erties of HBGs reflected obviously different Chen et al., Science 382, 559–565 (2023) 3 November 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Ionic transmission characteristics of nongated hydrogels and cascade- heterogated biphasic gels. (A) Homogeneous and (B) cross-interface ion transmission of the nongated hydrogel and HBGs, respectively. Scale bar, 3 mm. (C and D) Ionic current properties of the nongated hydrogels and (E and F) the HBGs containing 1 mM K+, Ca2+, or Fe3+ under varied applied voltages. (G) Statistically normalized ionic signal intensities of the HBGs with different phase compositions (data are mean ± SD, N = 3). (H) Normalized ionic signal intensities at 5 V of HBG-5 containing 1 mM of different ions. The ionic cross-interface transport capacity corresponds to the ionic hydration-dehydration energy (data are mean ± SD, N = 3). (I) Obvious distinction of the pulse-based K+, Ca2+, and Fe3+ ionic current properties of the HBG-5 under electrical signal stimulations with varying frequencies of 1, 5, and 10 Hz. (J) Calculated ionic signal intensity (S) ratio of K+/Ca2+ in the cascade- heterointerface system with different cascade degrees. The inset shows the equivalent circuit of n cross-interface ionic transmission units. trends under a range of varied applied volt- ages (Fig. 2, E and F). The voltage threshold of HBG-3 with 1 mM K+ was 5 mV, and 40 and 60 mV, respectively, with Ca2+ and Fe3+. To gain further insight into the relationship between the ionic cross-interface transmission and the phase-separation composition, ionic sig- nal transmission properties of HBGs varying with cascade-heterogated effects were character- ized. The ionic currents of the low-conductivity organogel separated from the biphasic gel pre- cursors were initially measured. All samples were in a nonconductive state at 5 V, confirm- ing that the ions were not stored and then continuously transported in the LC phase (fig. S15). The differences in ionic (K+, Ca2+, Fe3+) sig- nals derived from the heterojunction biphasic gel with a single and uniform heterointerface were still rough in comparison with HBGs fea- turing cascade-heterogated interfaces (figs. S16 and S17). By modulating cascade-heterogated effects of HBGs, the ratios of on-state transfer potentials and the normalized signal intensi- ties between different cations exhibited increas- ing trends induced by increasing the degrees of microphase separation (Fig. 2G and figs. S18 and S19A). Typically, the ratios of voltage thresholds (K+ to Fe3+ or K+ to Ca2+) of HBG-5 were 66.76 and 53.34. The normalized signal intensity ratios of K+ to Fe3+ and K+ to Ca2+ in HBG-5 reached 5162.40 and 2419.57. In con- trast, the values of K+ to Fe3+ and K+ to Ca2+ in nongated hydrogels were only 2.19 and 1.74, respectively. Such homogeneous ion transmis- sion depended on the ionic intrinsic transfer coefficients under an electric potential. The heterogated effect also held for iontronic sys- tems in the presence of a high cation concen- tration up to 100 mM (fig. S20). Otherwise, the normalized signal intensities and related volt- age thresholds of HBG-5 indicated that different ionic cross-interface transport capacities were related to the theoretical trends of their hydration- dehydration energies (Fig. 2H, fig. S19B, and table S1). For both isovalent and aliovalent ions, the ion with low hydration-dehydration energy corresponded to low cross-interface DE, thus exhibiting a relatively high ionic signal inten- sity and a low voltage threshold. Compared with the homogeneous ion trans- mission of nongated hydrogels (fig. S21), cascade- heterogated effects also demonstrated an obvious distinction between pulse-based K+, Ca2+, and Fe3+ ionic current properties under electrical stimulations at frequencies from 0.1 to 100 Hz (Fig. 2I and fig. S22). The simulated relationship between ion dynamics and pulse-based frequen- cies (0.1 to 1000 Hz) at the heterogated interface was conducted, in which the simulated trans- mitted ion number and normalized simulated current were unrelated to the stimulus frequency (fig. S23). The duration of the high level in a sin- gle pulse cycle far exceeded the expected time Chen et al., Science 382, 559–565 (2023) 3 November 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Electronic-to-multi-ionic signal transmission of cascade-heterogated biphasic gels. (A) Ionic current characteristics of the HBG-5 containing different ions. The inset showed the corresponding normalized signal intensities at 5 V (data are mean ± SD, N = 3). (B) Simulated signal intensity ratios of the HBGs with different mixtures of K+ and Ca2+ to that of the HBG with K+ (blue line), which were attributed to the different occupation proportions of K+ (Mmix K concentration ratio) and Ca2+ (Mmix Ca2þ as the Ca2+ molar concentration ratio) at the heterogated interface, and simulated independent signal intensity ratios of K+ (Smix K þ) þ as the K+ molar to that of Ca2+ (Smix Ca2þ) in the mixed system (orange line). (C) Schematic of preferential K+ transport and hysteretic Ca2+ transport at the heterogated interface. (D to F) Normalized conductance of the HBG-5 containing different ions (data are mean ± SD, N = 3). (G) Hierarchical ionic signal transmission. The pulse-transmitted ionic signals of the cascade-heterogated iontronics were successively dominated by K+, Ca2+, and Fe3+ against different electrical pulse signals from 5 to 500 mV (data are mean ± SD, N = 3). (H) Selective Ca2+ cross-stage transmission derived from the chemical-heterogated iontronics (data are mean ± SD, N = 3). Chen et al., Science 382, 559–565 (2023) 3 November 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Bioionic neurohumoral modulation. (A) Conceptual schematic of the iontronic signal transduction from electronic control circuits to bioionic neural circuits. (B) In an abiotic-biotic system, the cardiac electrical activities of a bullfrog heart were modulated by the bioionic neurohumoral signals from the iontronic transduction devices. Scale bars, 5 mm. (C) Exogenous bioionic neurohumoral signals modulate the permeability of cardiomyocytes and consequently alter the K+ outflux and the Ca2+ influx in the cell membrane. (D and E) Bioionic neurohumoral modulation in the cardiac electrical activities of bullfrog hearts based on the cascade- heterogated (D) or chemical-heterogated (E) iontronic transduction devices (data are mean ± SD, N = 3). Chen et al., Science 382, 559–565 (2023) 3 November 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E for the cross-interface transmission of an ion, indicating that the cross-interface ion trans- port of HBG systems would be frequency- independent (1 to 1000 Hz). To model cascade-heterogating effects, a series circuit model based on the nonlinear Poisson-Boltzmann theory was constructed to illustrate the relationship between cascaded heterointerfaces and K+/Ca2+ transport dis- tinction (Fig. 2J) (26, 27). The total energy con- sumption (º RTot) can be divided into n units. For one cross-interface unit, the energy con- sumption can be divided into four sections: the consumption of transport in the IE phase (RIE), across the heterogated interface from the IE phase to the LC phase (RIn), in the LC phase (RLC), and across the heterogated interface from the LC phase to the IE phase (ROut) (fig. S12B). RIE was negligible owing to the weak resistance in the IE phase; and ROut was also ignored because the energetically favorable reconfiguration for rehydration. On the basis of different cross-interface DE values, the ionic signal intensity (S) ratio of K+/Ca2+ (which was equivalent to the total transport resistance ra- tio of K+ to Ca2+) was expressed as SKþ =S Ca2þ ðnÞ ¼ 2 ¼ 2 (cid:1) nRLC Ca2þ nRLC Kþð (cid:3) (cid:1) þ nRIn Ca2þ Þþ nRIn Kþð Þ (cid:3) Þ ð RLC Ca2þ RLC Kþð Þ ð þ N RIn Ca2þ RIn Kþð 1 þ N Þ Þ ð1Þ where N ¼ RIn Kþð Þ Þ, and this N was used to com- RLC Kþð pare the inf luence between the heterogated interface and the LC phase on the ion trans- port. When N→∞, the system could be regarded as the ideal cascade-heterointerface limit (de- noted as the ideal HBG system) ð SKþ =S Ca2þ Cascade (cid:2) heterointerface limit (cid:4) ¼ 2exp (cid:1) DEIn Ca2þ Þ (cid:2) 2eDVIn Ca2þ (cid:2) DEIn Kþð (cid:3) (cid:1) kBT (cid:3) (cid:1) RIn Ca2þ RIn Kþð Þ (cid:5) þ eDVIn Kþð Þ Þ ¼ 2 (cid:3) ð2Þ In contrast, when N→0, the system could be regarded as a single interface (denoted as a heterojunction) SKþ =S ð Ca2þ Single interface Þ ¼ 2 (cid:3) (cid:1) RLC Ca2þ RLC Kþð Þ ð3Þ In practical conditions, the heterointerface density (º N) is often limited, and so the ionic signal intensity ratio of the HBG system cannot reach the theoretical limit. However, the cascade- heterogated interface of iontronics was suffi- ciently dense to ensure a high SKþ =S Ca2þ . Electronic-to-multi-ionic signal transmission of cascade-heterogated biphasic gels Figure 3A shows the anomalous decrease in the ionic current and the normalized signal intensity of HBG-5 containing a mixture of 1 mM K+ and 1 mM Ca2+. The presence of Ca2+ increased the total ion concentration, but the normalized signal intensity was only 0.025, which was one-fortieth that with K+ alone. This result can be attributed to the fact that K+ has a lower hydration-dehydration energy and a smaller ion-hydration size than Ca2+, and so it was transported more favorably across the heterogated interface to consecutively un- dergo partial dehydration and rehydration under the stimulation of an electrical signal, resulting in relatively efficient ion transmission. However, the Ca2+ cross-interface mobility re- quired a greater transfer energy, which reduced its own transport efficiency and formed inter- facial occupations, causing steric hindrance and electrostatic repulsion at the heterogated interface, thereby further hindering K+ and Ca2+ transport. Thus, the simulated signal in- tensity ratio of HBGs with different mixtures of K+ and Ca2+ to that of HBGs with only the same concentration of K+ was negatively linearly cor- related with the Ca2+ occupation ratio (Mmix Ca2+) at the heterogated interface (Fig. 3B, blue line). However, despite the decrease of ionic trans- mission paths resulting from such occupation, the simulated independent signal intensity ratio of K+/Ca2+ in the mixed system remained at greater than three orders of magnitude (as mix Ca2þ ≤ 0:91) (orange line). Figure 3C shows M preferential K+ transport and hysteretic Ca2+ transport DEKþ < Ca2þ Þ at the heterointer- ð face under segmental electrical signal stimula- tions. The Ca2+ occupation blocked the ion paths of the heterogated interface, also result- ing in a relatively low ionic conductance below ~3 V. As the applied electrical signal increased, the normalized conductance curve suddenly increased, which was mainly attributed to the increased transport of hysteretic Ca2+ (Fig. 3D). This effect was also confirmed in HBGs con- taining different binary ionic mixtures (Fig. 3, E and F, and figs. S25 and S26). DE Such cascade-heterogatings can fundamen- tally enlarge the hierarchy discrepancy of multi- ion transfer energy barriers for hierarchical signal transmission. The real-time variations in the transmitted ionic signal strengths derived from HBGs incorporated with ternary ionic mixtures were monitored by inductively coupled plasma mass spectrometry (fig. S27). Figure 3G and fig. S28 show that pulse-transmitted ionic signal strengths were successively dominated by K+, Ca2+, and Fe3+ against different input voltage signals ranging from 5 to 500 mV. In the absence of an input voltage signal, the HBGs maintained stable ion storage and avoided ion free diffusion to the surrounding aqueous phase environment. By applying a 5-mV pulse- based electrical signal with a frequency of 5 Hz, the transmitted K+ signal of the HBG was trig- gered, and the transport of Ca2+ and Fe3+ re- mained hysteretic. After 2 hours, the K+ signal strength increased from 6 to 400 parts per billion. Subsequently, by increasing the ap- plied voltages to 100 and 500 mV, Ca2+- and Fe3+-dominated signals were observed, respec- tively. The cascade-heterogated effect allowed for the preferential transport of multiple ions by the sorting of ionic transfer energy bar- riers. Meanwhile, such transmitted ionic signal strengths of HBGs have already achieved lev- els compatible with those of specific physio- logical bioactivities (28, 29). Unlike ion-selective membranes and hydrogels, our HBGs can store a stable ion source on demand and efficiently process intrinsic ionic signals to the ambient aqueous environment. The hydrogel iontronics inevitably led to uncontrolled ionic diffusion into the external environment, making it im- possible to modulate or regulate ion transmis- sion (fig. S29). Considering the advantageous transport or- der of one type of ion under specific condi- tions, the hierarchical ion transmission could be regarded as a form of ion selectivity. We also designed a chemical-heterogated architecture to further regulate the ionic transfer energy barrier hierarchy for ion-selective cross-stage transmission. Owing to the coordination ef- fects of specific ligand groups (18-crown-6 to K+ and carboxylic group to Fe3+) as chemical assistance in the IE phase (fig. S30A), ionic migration behaviors tended to be hopping between ligands to overcome dissociation en- ergy barriers (30, 31). Thus, the corresponding total energy consumption (º RTot) continued to increase, leading to a higher energy con- sumption in the IE phase than across the het- erogated interface and breaking the inherence in ionic transport hierarchies that originated from their intrinsic hydration-dehydration en- ergies. Such cascaded chemical-heterogated effects rendered the total equivalent transfer energy barriers of K+ and Fe3+ significantly greater than that of Ca2+, resulting in Ca2+ cross-stage signal and transmission restric- tions of the other signal (Fig. 3H and fig. S30B). In contrast, dual-ligand (18-crown-6 and car- boxylic group) hydrogel iontronics had no Ca2+ selectivity (fig. S31). In the homogeneous hydro- gel network, chemical ligand effects were unable to further establish ion-scale confined-gating properties that governed ion transport. Bioionic neurohumoral modulation Neuronal signaling relies on the controlled transmission of various neurotransmitter mol- ecules and ions (32, 33). For iontronics, the aim is to regulate multiple biofunctional ionic signal transductions to interface with and con- trol physiological activities in aqueous-based bioenvironments (34). To that end, the cascade- heterogated iontronic transduction device devel- oped in this study to convert electronic input signals into diverse bioionic signals demon- strates its potential for application as a biocom- munication carrier (Fig. 4, A and B). Meanwhile, Chen et al., Science 382, 559–565 (2023) 3 November 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E the good biocompatibility of the developed sys- tem was also demonstrated (figs. S32 and S33). Figure 4, C and D, outlines the transduction of biofunctional neurohumoral signals by cascade- heterogated and chemical-heterogated devices to regulate the permeability of cardiomyocytes and modulate cardiac electrical activities in bullfrog hearts. These results were confirmed by the alterations in the electrocardiogram (figs. S34 and S35 and movies S1 and S2). For abiotic-biotic systems, we anticipate that such iontronics could act as bridges for biocompa- tible signal processing and transmission between electronic control circuits and bioionic neural circuits. Conclusions Here we reported cascade-heterogated iontronics for diverse ionic signal transmission. These gels integrated opposite binary phase structures to form cascaded ionic transfer free energy barriers that were highly correlated with the hydration- dehydration energies of different ions. More impressively, cascade-heterogated interfaces fundamentally enlarged and controlled the hierarchical discrepancy of ionic transfer en- ergy barriers to achieve multi-ionic hierarchi- cal transmission and ion-selective cross-stage transmission. Until now, ion-gating has been primarily manifested in one- or two-dimensional systems; in this work, we developed a cascad- ing strategy to expand the ion-gating mecha- nism to three-dimensional intronic systems for electronic-to-multi-ion transmission. Most iontronic and electronic devices employing con- ventional ion-gating or nongating struggle to effectively process multi-ion signal carriers in their potential applications (tables S2 and S3); however, these properties are critical for inter- facing with and matching complex biological systems. HBG-based iontronics offer not only multispecies bioionic signals that are highly com- patible with aqueous-phase biological systems but also advance programmable signal trans- mission functionalities that facilitate the parallel and multiplexed transmission of various bio- signals in abiotic-biotic systems. These cascade- heterogated iontronics serve as diverse platforms for possible extensions to more-biofunctional car- riers and other signal transduction mechanisms. Meanwhile, it would be worthwhile to study the weight and coupling features of multispecies sig- nal interchange in these cascade-gating systems. Despite the long road ahead for the construction of high-performance artificial neural networks, we expect that such cascade-heterogated net- works will be applicable in hardware imple- mentation for neuromorphic gating features, while also providing more signal morpholo- gies and processing. RE FERENCES AND NOTES 1. J. Chen, Y. Kanai, N. J. Cowan, N. Hirokawa, Nature 360, 674–677 (1992). 2. T. Yoshimura et al., Cell 120, 137–149 (2005). 3. X. Xia, C. M. Spadaccini, J. R. Greer, Nat. Rev. Mater. 7, 683–701 (2022). 4. Y. Liu et al., Nat. Biotechnol. 38, 1031–1036 (2020). 5. S. Zhao et al., Adv. Mater. 30, e1800598 (2018). 6. G. T. Go, Y. Lee, D. G. Seo, T. W. Lee, Adv. Mater. 34, e2201864 (2022). 7. Y. Kim et al., Science 360, 998–1003 (2018). 8. Z. Nie, J. W. Kwak, M. Han, J. A. Rogers, Adv. Mater. 10.1002/adma.202205609 (2022). 9. Y. Lee et al., Nat. Biomed. Eng. 7, 511–519 (2023). 10. P. C. Harikesh et al., Nat. Mater. 22, 242–248 (2023). 11. T. Sarkar et al., Nat. Electron. 5, 774–783 (2022). 12. T. Wang et al., Nat. Electron. 5, 586–595 (2022). 13. W. Wang et al., Science 380, 735–742 (2023). 14. S. Chun et al., Nat. Electron. 4, 429–438 (2021). 15. S. H. Jo et al., Nano Lett. 10, 1297–1301 (2010). 16. L. Q. Zhu, C. J. Wan, L. Q. Guo, Y. Shi, Q. Wan, Nat. Commun. 5, 3158 (2014). 17. H. Wei et al., Nat. Commun. 12, 1068 (2021). 18. P. C. Harikesh et al., Nat. Commun. 13, 901 (2022). 19. C. E. G. Hoskin, V. R. Schild, J. Vinals, H. Bayley, Nat. Chem. 14, 650–657 (2022). 20. Y. Dobashi et al., Science 376, 502–507 (2022). 21. P. Wang, G. Li, W. Yu, C. Meng, S. Guo, Adv. Electron. Mater. 8, 2101269 (2022). 22. X. Liu, J. Liu, S. Lin, X. Zhao, Mater. Today 36, 102–124 (2020). 23. R. Epsztein, R. M. DuChanois, C. L. Ritt, A. Noy, M. Elimelech, Nat. Nanotechnol. 15, 426–436 (2020). 24. C. Lu et al., J. Am. Chem. Soc. 143, 14242–14252 (2021). 25. P. A. Fernandes, M. N. D. S. Cordeiro, J. A. N. F. Gomes, J. Phys. Chem. B 104, 2278–2286 (2000). 26. T. S. Badessa, V. A. Shaposhnik, Electrochim. Acta 231, 453–459 (2017). 27. N. Rajeswari, S. Selvasekarapandian, M. Prabu, S. Karthikeyan, C. Sanjeeviraja, Bull. Mater. Sci. 36, 333–339 (2013). 28. E. K. Hoffmann, I. H. Lambert, S. F. Pedersen, Physiol. Rev. 89, 193–277 (2009). 29. M. Simard, M. Nedergaard, Neuroscience 129, 877–896 (2004). 30. W. Wei, C. Xu, J. Ren, B. Xu, X. Qu, Chem. Commun. 48, 1284–1286 (2012). 31. P. Lin, S. Ma, X. Wang, F. Zhou, Adv. Mater. 27, 2054–2059 (2015). 32. S. Patel, R. Docampo, Trends Cell Biol. 20, 277–286 (2010). 33. S. T. Keene et al., Nat. Mater. 19, 969–973 (2020). 34. Y. Hou, X. Hou, Science 373, 628–629 (2021). 35. W. Chen et al., Data for: Cascade-heterogated biphasic-gel iontronics for electronic-to-multi-ionic signal transmission processing, version 1, Zenodo (2023); https://doi.org/ 10.5281/zenodo.8319623. AC KNOWLED GME NTS All animals used in the experiment were approved by the Animal Protection Ethics Committee of Capital Medical University (Ethics Number: AEEI-2014-062). We are grateful to X. Dong (Experimental Center for basic medical teaching, Basic Medical Sciences, Capital Medical University) for assistance with the biological experiments, and K. Wang (Capital Medical University) for helpful discussions. Funding: This research was supported by the National Natural Science Foundation (21988102, 22275183, 22122207, 11825203), the National Key R&D Program of China (2022YFB3805904, 2022YFB3805900), the CAS Project for Young Scientists in Basic Research (YSBR-039), and the China Postdoctoral Science Foundation (2022M713226). Author contributions: W.C., L.Z., S.Z., and Z.Z. contributed equally to this work. Z.Z., L.W., and W.C. proposed the research direction. Z.Z., L.W., and L.J. guided the project. W.C. and Z.Z. co-designed and performed the experiments. L.Z. and Z.X. performed the theoretical calculations. S.Z. and H.L. performed the biology experiments. Z.Z. and W.C. drafted the manuscript. All authors joined the data discussion and revised the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the main text or the supplementary materials. The data that support the findings of this study are available at Zenodo (35). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg0059 Materials and Methods Supplementary Text Figs. S1 to S35 Tables S1 to S4 References (36–72) Movies S1 and S2 Submitted 26 November 2022; resubmitted 22 July 2023 Accepted 22 September 2023 10.1126/science.adg0059 Chen et al., Science 382, 559–565 (2023) 3 November 2023 7 of 7
10.1126_science.adg9232
RES EARCH ORGANOMETALLICS Oxidative addition of an alkyl halide to form a stable Cu(III) product Yongrui Luo1†, Yuli Li1†, Jian Wu1, Xiao-Song Xue1*, John F. Hartwig2*, Qilong Shen1* The step that cleaves the carbon-halogen bond in copper-catalyzed cross-coupling reactions remains ill defined because of the multiple redox manifolds available to copper and the instability of the high- valent copper product formed. We report the oxidative addition of a-haloacetonitrile to ionic and neutral copper(I) complexes to form previously elusive but here fully characterized copper(III) complexes. The stability of these complexes stems from the strong Cu−CF3 bond and the high barrier for C(CF3)−C(CH2CN) bond-forming reductive elimination. The mechanistic studies we performed suggest that oxidative addition to ionic and neutral copper(I) complexes proceeds by means of two different pathways: an SN2-type substitution to the ionic complex and a halogen-atom transfer to the neutral complex. We observed a pronounced ligand acceleration of the oxidative addition, which correlates with that observed in the copper-catalyzed couplings of azoles, amines, or alkynes with alkyl electrophiles. C opper-mediated cross-coupling reactions have become some of the most powerful methods for the construction of carbon- carbon (C−C) and C−heteroatom bonds (1–4). Early efforts in this field focused mainly on the coupling of sp2-hybridized car- bon electrophiles, and in recent years, research has expanded to encompass the mild coupling of sp3-hybridized carbon electrophiles (5, 6). Much progress has been made recently on copper-mediated or copper-catalyzed alkynyla- tion (7–9), alkylation (10–12), arylation (13–15), and amination (16, 17) of alkyl halides, pro- viding an alternative and practical method for the installation of functional groups on alkyl chains and rings. Despite this expansion of the scope of the cross-coupling reactions of alkyl electrophiles that use copper, the mechanism of these reactions is poorly understood. Previously, two distinct cycles were proposed for copper-catalyzed cross-coupling reactions of alkyl electrophiles. One cycle comprises a two-electron Cu(I)/Cu(III) manifold, and one comprises a stepwise Cu(I)/Cu(II) manifold involving initial single-electron transfer (SET) between the Cu(I) center and the alkyl electro- phile to generate a Cu(II) intermediate and an alkyl radical, followed by transfer of the func- tional group from the resulting Cu(II) species to the alkyl radical (18–22) (Fig. 1A). Differen- tiation of these two pathways is challenging because the higher-valent copper intermedi- ates in the reactions, particularly the putative Cu(III) intermediates, are highly reactive and typically elude detection (23–26). 1Key Laboratory of Organofluorine Chemistry, Center for Excellence in Molecular Synthesis, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032, PR China. 2Department of Chemistry, University of California, Berkeley, Berkeley, CA 94720, USA. *Corresponding author. Email: xuexs@sioc.ac.cn (X-S.X.); jhartwig@berkeley.edu (J.F.H.); shenql@sioc.ac.cn (Q.S.) †These authors contributed equally to this work. Consequently, isolable, well-defined alkyl Cu (III) complexes are rare. In the early 2000s, seminal work from the groups of Bertz, Ogle (27–29), and Gschwind (30) demonstrated that reactions of alkyl iodides with ionic Gilman reagents generated Cu(III) species, which were characterized at a temperature below −93°C with rapid-injection nuclear magnetic reso- nance (RI-NMR) spectroscopy techniques (Fig. 1C). Yet, formation of the Cu(III) inter- mediates, even at this temperature, is fast, rendering studies on the mechanism of the reaction of alkyl halides with the Cu(I) species challenging. To probe whether a Cu(III) intermediate could be generated from the reaction of an alkyl halide with a Cu(I) species and to deter- mine how such an intermediate would form from an alkyl electrophile in a copper-mediated or copper-catalyzed cross-coupling, we rea- soned that two requirements must be met: (i) An alkyl electrophile with a high reduc- tion potential must be used so that forma- tion of the Cu(III) species from the starting Cu (I) species is thermodynamically favorable, and (ii) the barrier for reductive elimination from the Cu(III) intermediate must be higher than that for the oxidative addition that forms the Cu(III) species (Fig. 1B). The electron-withdrawing, strong-field tri- fluoromethyl ligand is known to stabilize both Cu(I) and high-valent Cu(III) metal centers as a result of p-back donation of electron density from the d orbitals on copper to the antibond- ing (s*) orbitals of the C–F group or the con- tracted bonding (s) orbitals on copper because of the high electronegativity of the fluorine (31). In the past three decades, a variety of well-defined trifluoromethyl Cu(III) complexes have been reported (26). Recent studies on the reactivities of these Cu(III) complexes showed that the barrier for reductive elimi- nation to form a C(sp3)−CF3 bond from either the ionic Cu(III) complex [Cu(CF3)3(alkyl)]− or the five-coordinate neutral Cu(III) complex [(bpy)Cu(CF3)2(Me)] is high (32–34). In addi- tion, stable, well-characterized trifluoromethyl Cu(I) species—including [CuCF3] (35), which contains no additional ligands; [(Phen)CuCF3] (36), which contains a dinitrogen-donor ligand; [(NHC)CuCF3] (37), which contains an N- heterocyclic carbene (NHC) ligand; and [Cu(CF3)2]− (38, 39), which is an ionic Cu(I) cuprate—were reported to react with aryl halides to give tri- fluoromethylarene products in good yields. In light of this prior work, we proposed that an appropriate trade-off between reactivity and stability of trifluoromethyl Cu(I) and Cu(III) complexes could meet the aforementioned re- quirements and provide an opportunity to in- vestigate the mechanism of the reaction of alkyl halides with Cu(I) species to form stable Cu(III) products. On the basis of the above rationale, we studied the reactions of alkyl halides with trifluoromethyl-Cu(I) complexes (Fig. 1). Stable [Ph4P]+[Cu(CF3)2]− and [(bpy)Cu(CF3)] (bpy, bipyridine) were chosen initially as the Cu(I) complexes that represent the ligandless “ate”- type Cu(I) complex and the neutral bipyridine- ligated Cu(I) complex, respectively. These complexes would allow us to probe the dif- ferences between the reactivity of two differ- ent types of Cu(I) complexes and the effect of the nitrogen ligand on the oxidative addition process. We chose a-haloacetonitrile XCH2CN [in which X (halogen) is Cl, Br, or I] as the alkyl electrophile because XCH2CN is more electro- philic than other alkyl halides, reducing the barrier for oxidative addition to Cu(I). In ad- dition, because of the electron-withdrawing property of the cyano group, the barrier for C−C bond-forming reductive elimination from a [(ligand)CuIII(CF3)2(CH2CN)] complex could be sufficiently high for the product to be ob- served directly and possibly isolated. We re- port the isolation of Cu(III) complexes from oxidative addition of haloacetonitrile to ionic and neutral trifluoromethyl Cu(I) complexes. Mechanistic investigation of these reactions, conducted with a combination of computa- tional and experimental studies, shows that anionic and neutral complexes react with the same alkyl halide by means of distinct mechanisms. Isolation and characterization of Cu(III) products We initially studied the reactions of [Ph4P]+ [Cu(CF3)2]− (1a) or a neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) with haloacetonitriles ClCH2CN (2-Cl), BrCH2CN (2-Br), and ICH2CN (2-I) or alkyl tosylate TsOCH2CN (2-OTs) (Fig. 2A). The reaction of 2.0 equivalents of anion 1a and bromide 2-Br in dimethyl sulfoxide (DMSO) occurred smoothly at room temperature after 3.0 hours to give two pre- viously unknown species in an approximate Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Mechanism of Cu-mediated cross-coupling. (A) General mechanism for Cu-catalyzed cross-coupling reaction. (B) Strategy that inverted the barrier for oxidative addition (OA) and reductive elimination (RE) in the catalytic cycle for Cu-catalyzed cross-coupling. (C) State-of-the-art observations of oxidative addition of alkyl halide to Cu(I) species. M, metal or quasi-metal; X, halogen; Y, dummy ligand; TM, transmetalation. 1:1 ratio in quantitative yield, as determined by 19F NMR spectroscopy. One species, cor- responding to chemical shifts of −33.4 and −34.4 parts per million (ppm) in a 1:2 integral ratio in the 19F NMR spectrum, was assigned as the tetracoordinate ionic Cu(III) complex [Ph4P]+[Cu(CF3)3(CH2CN)]− (3a) (fig. S1). Cu(III) complex 3a was stable enough to be isolated and characterized by 1H, 19F, and 31P NMR spectroscopies and elemental analysis. The structure of complex 3a was further con- firmed by single-crystal x-ray diffraction, which revealed a typical square-planar geometry (Fig. 2C, top). The second complex, corresponding to a chemical shift at −27.0 ppm in the 19F NMR spectrum, was assigned as a cuprate(I) spe- cies, [Ph4P]+[Cu(CF3)(Br)]− (1c). We presume that complexes 3a and 1c were generated from transmetalation of the oxidative addi- tion product [Ph4P]+[Cu(CF3)2(CH2CN)(Br)]− with Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a). It was found that complex 1c was much less reactive than complex 1a. It reacted with 1.0 equivalent of bromide 2-Br in DMSO, result- ing in 30% conversion and 9% yield of 3a after 3 hours at room temperature. Thus, the for- mation of bromocuprate 1c does not affect the oxidative addition of 2-Br to [Ph4P]+[Cu(CF3)2]− (1a). As did the reaction of bromide 2-Br, the reaction of iodide 2-I with cuprate 1a occur- red smoothly to give Cu(III) complex 3a in more than 90% yield and [Ph4P]+[Cu(CF3)(I)]− (1d) in 62% yield, respectively. By contrast, the reaction of chloride 2-Cl with cuprate 1a was much slower, and the starting materials re- mained intact even after 5 hours at room tem- perature (Fig. 2A). We also studied the reaction of tosylate 2-OTs with 1a and found that com- plex 1a was fully converted within 3 hours at 25°C, but the yield for the formation of Cu(III) complex 3a was much lower (15%) than those from the reactions of 2-Br or 2-I (Fig. 2A). The reaction of neutral Cu(I) complex [(bpy) Cu(CF3)] (1b) with BrCH2CN (2-Br) occurred much faster than that of ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a). The reaction of 1b with bromide 2-Br in N,N-dimethylformamide (DMF) generated trans-[(bpy)Cu(CF3)2(CH2CN)] (trans-4) and cis-[(bpy)Cu(CF3)2(CH2CN)] (cis- 4) in 56 and 4% yields, as well as [(bpy)CuBr], respectively, after just 1 min at room temper- ature (Fig. 2B). Complex trans-4 was fully characterized by 1H and 19F NMR spectros- copies, as well as elemental analysis. X-ray diffraction of the single crystal of trans-4 shows that it adopts a distorted square- pyramidal geometry. The H2C−Cu−N angle to the apical nitrogen is 121.9°, and that to the equatorial nitrogen is 160.2°; the length of the N−Cu bond at the apical position (2.236 Å) is considerably longer than that of the N−Cu bond in the equatorial position (1.987 Å), indi- cating that the geometry is closer to a square- based pyramid than a trigonal bipyramid (Fig. 2C, bottom). Reaction of complex 1b with iodide 2-I also occurred quickly to give trans-4 and cis-4 in 67 and 11% yields, respectively, under similar conditions (Fig. 2B). In contrast to cuprate 1a, the neutral complex 1b reacted with chloride 2-Cl to give trans-4 in 49% yield at room temperature after just 1 hour (Fig. 2B). These rapid rates indicate that the neutral Cu(I) com- plex 1b is much more reactive toward the alkyl halide than is ionic Cu(I) complex 1a, reflect- ing a sizable ligand acceleration (Fig. 2D). Such a phenomenon has previously been ob- served in various copper-mediated or -catalyzed cross-coupling reactions (40, 41). The large difference in reactivity of ionic and neutral Cu(I) complexes led us to conduct Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E reactions that would reveal the mechanisms by which the two complexes react with alkyl halides. The reaction of bromide 2-Br with ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) was sensitive to the polarity of the solvent. The reactions conducted in polar solvents, such as DMSO, N,N-dimethylacetamide (DMAc), and DMF, proceeded to 96, 87, and 76% con- version at room temperature over 5 hours and afforded Cu(III) complex 3a in 93, 70, and 60% yields, respectively (table S1), whereas the same reactions in less-polar solvents, such as tetrahydrofuran (THF) or CH2Cl2, occurred more slowly (24 and 12% conversions after 5 hours at room temperature) to give com- plex 3a in just 13 and 11% yields, respectively (Fig. 2E). By contrast, the reaction of chloride 2-Cl or bromide 2-Br with the neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) was not sensitive to the polarity of the solvent. Reactions of com- plex 1b with chloride 2-Cl in the more-polar solvent DMF or the less-polar solvent THF oc- curred to full conversion after 1 hour at room temperature to afford Cu(III) complex trans-4 in 59 and 82% yields, respectively (Fig. 2E). Quantitative assessment of the reaction of chlo- ride 2-Cl with complex 1b in THF and DMF at −5°C showed that the rates of the two reac- tions are similar [(3.02 ± 0.10) × 10−3 M−1·s−1 in THF and (3.43 ± 0.32) × 10−3 M−1·s−1 in DMF, Fig. 2. Oxidative addition of alkyl halides to Cu(I) complexes. (A) Oxidative addition of XCH(R)CN with ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a). (B) Oxidative addition of XCH(R)CN with [(bpy)Cu(CF3)] (1b). (C) ORTEP (Oak Ridge Thermal Ellipsoid Plot) diagrams of [Ph4P]+[Cu(CF3)3(CH2CN)]− (3a) (countercation Ph4P+ is omitted for clarity) and trans-[(bpy)Cu(CF3)2(CH2CN)] (trans-4). Ellipsoids are shown at the 50% level. (D) Reaction progress for ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) with BrCH2CN (2-Br) at 298 K (blue) or neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) with BrCH2CN (2-Br) at 298 K (red). (E) Solvent effect on the reactions of BrCH2CN with [Ph4P]+[Cu(CF3)2]− (1a) (blue) or [(bpy)Cu(CF3)] (1b) (red). DCM, dichloromethane; MeCN, acetonitrile (methyl cyanide). Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E respectively (fig. S25)]. The different sensitivities of the ionic and neutral Cu(I) species to sol- vent polarity indicate that these Cu(I) com- plexes might react with the alkyl halide through different mechanisms. Kinetic studies To investigate the effect of the properties of the C−X bond of the haloacetonitriles on the reactions with cuprate [Ph4P]+[Cu(CF3)2]− (1a) or neutral [(bpy)Cu(CF3)] (1b), we compared the rates of reactions of 1a or 1b with XCH2CN quantitatively by monitoring the 19F NMR sig- nals corresponding to 1a for reaction of com- plex 1a and the signals corresponding to trans-4 and cis-4 for the reaction of complex 1b. These studies showed that reactions of the ate complex are first order in complex 1a and first order in bromide 2-Br (Fig. 3A and figs. S3 to S6). The reaction of complex 1a with iodide 2-I [(8.54 ± 0.04) × 10−3 M−1·s−1] was about five times as fast as that with bromide 2-Br [(1.78 ± 0.03) × 10−3 M−1·s−1 at 25°C]. Because complex 1a reacted with chloride 2-Cl slowly at room temperature (Figs. 2A and 3A), we studied the reaction at elevated temperature. At 60°C, the reaction occurred with a rate constant of (1.99 ± 0.14) × 10−4 M−1·s−1. The rate constant for the reaction of complex 1a with bromide 2-Br at 60°C was estimated to be 4.05 × 10−2 M−1·s−1 on the basis of the Eyring analysis in Fig. 3C, which is roughly 200 times as fast as that with chloride 2-Cl. The rates of the reactions of the halides with neutral complex 1b were measured with 19F NMR spectroscopy below room temperature because of their high rate. The reaction of bro- mide 2-Br with [(bpy)Cu(CF3)] (1b) at −30°C proceeded to full conversion after 20 min. Kinetic studies showed that this reaction is first order in both reactants and that the rate constant at −30°C is (2.63 ± 0.05) × 10−2 M−1·s−1. At this temperature after 5 min, the reaction of chloride 2-Cl with [(bpy)Cu(CF3)] (1b) pro- ceeded to <1% conversion (Fig. 3B). However, the reaction of chloride 2-Cl with complex 1b at −5°C occurred with a rate constant of (3.43 ± 0.32) × 10−3 M−1·s−1. According to the Eyring analysis shown in Fig. 3C, the rate constant for reaction of complex 1b with bromide 2-Br at −5°C was estimated to be 0.37 M−1·s−1, which is roughly 100 times greater than that of the reaction of complex 1b with chloride 2-Cl. The rate dependence of these reactions of the ionic and neutral Cu(I) species on the strength of the C−X bond and on the leaving group ability of the alkyl halides is consistent with typical Cu(I)- catalyzed cross-coupling reactions. To evaluate the effect of the steric proper- ties of the alkyl bromide on the reaction of the Cu(I) complex, we studied the reaction of Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) with the sec- ondary alkyl halide, 2-bromopropionitrile 2-Br-Me, and the tertiary alkyl halide, 2-methyl- 2-bromopropionitrile 2-Br-Me2 (Fig. 2A). Re- action of cuprate 1a with 2-bromopropionitrile 2-Br-Me occurred smoothly at room temper- ature over 5 hours to give Cu(III) complex 3b in 81% yield. Yet, this reaction of the more steri- cally hindered 2-bromopropionitrile 2-Br-Me was slower than that of the less-hindered bromo- acetonitrile 2-Br [(1.15 ± 0.01) × 10−3 M−1·s−1 versus (1.78 ± 0.03) × 10−3 M−1· s−1]. To de- termine whether the reactions of complex 1a occurred with inversion of configuration, we conducted the reaction of 1a with optically active (R)-2-bromopropionitrile (R)-2-Br-Me, but the enantiomers of the resulting product Fig. 3. Kinetic analysis. Kinetic data were fit to the expression of [1a]t = [1a]0e−kobs + c for (A) and [4]t = A − Be−kobs for (B), in which t is time and kobs are the apparent (observed) rate constants (pages S12 and S35 to S38 provide details about derivation of expression of corrected rate constant kcorr from kobs). (A) Kinetic profiles of oxidative addition of XCH(R)CN (where X is Cl, Br, or I; and R is H or Me) with ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) at 298 K. (B) Kinetic profiles of oxidative addition of XCH2CN (where X is Cl or Br) with [(bpy)Cu(CF3)] (1b) at 243 K. (C) Eyring analysis of the temperature dependence of the rate constants of oxidative addition of BrCH2CN with [Ph4P]+[Cu(CF3)2]− (1a) (blue), oxidative addition of BrCH2CN with [(bpy)Cu(CF3)] (1b) (green), and oxidative addition of ClCH2CN with [(bpy)Cu(CF3)] (1b) (red). (D) Effect of added free bipyridine on oxidative addition of ClCH2CN with [(bpy)Cu(CF3)] (1b) at 268 K. Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Five proposed pathways for the oxidative addition of haloacetoni- triles to ionic or neutral Cu(I) complexes and free energies of each species computed with DFT. The calculated activation free energies for the oxidative addition of BrCH2CN(2-Br) to the ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) in DMSO are given in blue, and those for oxidative addition of ClCH2CN (2-Cl) to the neutral [(bpy)Cu(CF3)] (1b) in DMF are given in red. The energies are in kilocalories per mole and indicate the relative free energies calculated at the PBE0-D3(BJ)/Def2-TZVP(SMD, solvent)//PBE0-D3(BJ)/Def2-SVP(SMD,solvent) level. SMD, solvation model density. 3b did not separate through chiral high- performance liquid chromatography (HPLC) column and only weakly absorbed in the ultra- violet, preventing assessment of the configu- ration of the product with chromatography or spectroscopy. The reaction of 1a with the ter- tiary alkyl halide, 2-methyl-2-bromopropionitrile 2-Br-Me2, did not afford the corresponding Cu(III) complex from oxidative addition. In- stead, Cu(III) complex [Ph4P]+[Cu(CF3)4]− and Cu(I) complex [Ph4P]+[Cu(CF3)(Br)]− formed in 31 and 7% yields, respectively, as determined by 19F NMR spectroscopy of the reaction mix- ture, as well as a few unidentified Cu(II) spe- cies, which were indicated by the color of the reaction mixture turning green. Analysis of the reaction mixture showed that isobutyronitrile 6 also formed through hydrodehalogenation. This observation suggests that the reaction forms the isobutyronitrile radical, which is too sterically hindered to combine with the trifluo- romethyl Cu(II) species and, instead, abstracts a hydrogen atom from the solvent. The reac- tion of 2-chloropropionitrile 2-Cl-Me with the neutral complex [(bpy)Cu(CF3)] (1b) occurred over 1 hour at 25°C to generate the correspond- ing Cu(III) complex trans-4b in 12% yield. This complex was characterized with 19F NMR spectroscopy because it was too unstable to be isolated. Inhibition studies To probe whether the oxidative additions of XCH2CN to Cu(I) complexes [Ph4P]+[Cu(CF3)2]− (1a) and [(bpy)Cu(CF3)] (1b) occur by means of a radical intermediate and whether a po- tential radical would form through initial SET, we first studied the reaction in the presence of radical inhibitor 2,2,6,6-tetramethylpiperidine- 1-oxyl (TEMPO) and in the presence of SET inhibitor 1,4-dinitrobenzene. The reaction of 1a with BrCH2CN 2-Br in the presence of 2.0 equiv- alents of TEMPO was complete after 3 hours at room temperature. The yield of this reaction (64%) was only 29% less than that in the ab- sence of TEMPO, and only 19% of the radical adduct TEMPO−CH2CN 5 was isolated (Fig. 2A). By contrast, TEMPO had a strong effect on the reaction of 1b with ClCH2CN 2-Cl, resulting in the formation of trans-4 in 10% yield and TEMPO-CH2CN 5 in 75% yield (Fig. 2B). SET inhibitor 1,4-dinitrobenzene did not noticeably affect the reaction of haloacetonitrile with either the ionic Cu(I) complex 1a or the neutral Cu(I) complex 1b. These results suggest that alkyl radi- cals are generated in the oxidative addition of 1a or 1b with bromoacetonitrile but that a SET pro- cess is unlikely to lead to the radical in either case. Effect of ligand The bipyridine ligand in complex 1b could in principle dissociate from the metal center to create a less sterically hindered intermediate that reacts with the alkyl halide. If reaction of ClCH2CN with neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) occurred through reversible dissociation Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E of the bipyridine, addition of free bipyridine to the reaction should substantially reduce the rate. The reactions in the presence of varying amounts of bipyridine (3.0, 6.0, and 10.0 equiv- alents versus 1b) occurred with nearly equal rate constants [(2.81 ± 0.01) × 10−3 M−1·s−1, (3.33 ± 0.21) × 10−3 M−1·s−1, and (2.91 ± 0.12) × 10−3 M−1·s−1, respectively] and were only slight- ly slower than the reaction in the absence of added 2,2′-bipyridine [(3.61 ± 0.08) × 10−3 M−1·s−1] (Fig. 3D). These data imply that re- versible dissociation of the ligand does not precede rate-limiting oxidative addition to 1b. Activation parameters To determine the enthalpy and entropy of ac- tivation of both reactions, we studied the ef- fect of the temperature on the rates. An Eyring plot of ln(k/T) versus 1/T (where k is the rate constant and T is temperature) for the reac- tion of ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) with bromide 2-Br between 25° and 37°C revealed an activation enthalpy, DH‡, of 17.7 ± 0.4 kcal/mol and an activation entropy, DS‡, of −12 ± 1 entropy units (e.u.) (Fig. 3C, blue). Likewise, an Eyring analysis of the reaction of neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) with 2-Br over a temperature range of −30° to −20°C revealed an activation enthalpy, DH‡, of 13.0 ± 0.6 kcal/mol and a similar activation entropy, DS‡, of −12 ± 2 e.u. (Fig. 3C, green). The activation entropies of the reactions of 1a and of 1b with 2-Br were similar, but the ac- tivation enthalpy for the reaction of 2-Br with 1b was much less than that with 1a. An Eyring analysis of the reaction of [(bpy)Cu(CF3)] (1b) with ClCH2CN (2-Cl) between −8° and 4°C revealed a DH‡ of 15.2 ± 0.8 kcal/mol and a DS‡ of −13 ± 3 e.u. (Fig. 3C, red), and this value for DH‡ is similar to, or even slightly less than, that for oxidative addition of BrCH2CN (2-Br) to cuprate(I) complex Ph4P+[Cu(CF3)2]− (1a) (17.7 ± 0.4 kcal/mol), even though the C−Cl bond in ClCH2CN (2-Cl) (70.5 kcal/mol) is much stronger than the C−Br bond in BrCH2CN (2-Br) (56.8 kcal/mol) and the chloride is less polarizable and is a poorer leaving group than bromide (42). To define the mechanism of the reaction of haloacetonitrile with the ionic or neutral Cu(I) species further, we performed density func- tional theory (DFT) calculations with the PBE0- D3(BJ) functional. On the basis of previous proposals for the mechanism of copper-mediated cross-coupling reactions (18–22) and the ex- perimental results reported in this work, we proposed five different pathways that could account for the oxidative addition of haloace- tonitrile to ionic and neutral Cu(I) complexes (Fig. 4). The first pathway (pathway A) in- volves an SN2 step and is a common type of two-electron oxidative addition of alkyl halides to late transition metals, such as Pt (43) and Pd (44). In addition, such a pathway was pro- posed by Whitesides (45) and Pearson (46) for the reaction of lithium dialkylcuprates with alkyl halides (the Corey-Posner reaction) to generate a Cu(III) intermediate, which would then undergo fast reductive elimination to form the C−C bond in the product. In our case, reaction of 1a through this mechanism would form Cu(III) species [CuIII(CF3)2(CH2CN)] or [Ph4P]+[CuIII(CF3)2(X)(CH2CN)]–, which would then undergo transmetalation with 1a to gen- erate [Ph4P]+[CuIII(CF3)3(CH2CN)]– (3a) and [Cu(CF3)X]–. Likewise, reaction with 1b would give [(bpy)CuIII(CF3)(CH2CN)]+ or [(bpy)CuIII (CF3)(X)(CH2CN)], which would undergo trans- metalation with 1b to give [(bpy)CuIII(CF3)2 (CH2CN)] (trans-4 or cis-4) and [(bpy)CuX], respectively (details about transmetalation are provided in fig. S31). A second two-electron mechanism for oxidative addition would be a concerted addition of the C−X bond to the metal center (pathway B), and this step would be the reverse reaction of concerted reductive elimination from a high-valent Cu center. A third pathway for oxidative addition of alkyl halides to a Cu(I) center could occur through consecutive single-electron steps. One commonly proposed mechanism for oxidative addition through single-electron steps is outer-sphere SET (OSET; pathway C), which dominates the redox manifolds of first-row transition metals, including Fe (47) and Ni (48). A fourth path- way and an alternative mechanism for oxida- tive addition through single-electron steps is halogen-atom transfer (XAT; pathway D) (49, 50). Ligated Cu(I) complexes are widely used in atom-transfer radical addition or polymeriza- tion (ATRA or ATRP) processes in which a XAT to Cu is involved (51). A fifth pathway, oxidative addition of the alkyl halide to Cu(I), could occur by initial ligand dissociation to generate a neutral, ligandless [CuCF3], which would undergo oxidative addition of the alkyl halide to generate a Cu(III) intermediate that would recoordinate the dative ligand and undergo transmetalation to give the final Cu(III) complex (pathway E) (52). The energy param- eters for these pathways were computed and are shown in Fig. 4. In all five proposed mech- anistic pathways, the formation of final product 3a from the reaction of 1a and the formation of trans-4/cis-4 from reaction of 1b involve a transmetalation step after initial formation of a Cu(III) complex. The absence of observed in- termediates before transmetalation and first- order rate behavior in the Cu(I) species (1a or 1b) rule out both rate-limiting transmetala- tion for these reactions and the initial gen- eration of the Cu(III) intermediate through reaction of two copper complexes. Evidence for XAT and SN2 pathways On the basis of the experimental results and DFT calculations, we concluded that reaction of the ionic Cu(I) complex [Ph4P]+[Cu(CF3)2]− (1a) with BrCH2CN (2-Br) likely proceeds by means of two different pathways, specifically the major fraction through an SN2-type pro- cess (pathway A) and the minor fraction through a XAT process (pathway D). We also con- clude that the reaction of the neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) with ClCH2CN (2-Cl) likely proceeds exclusively through a XAT process (pathway D). The following analy- sis of our experimental and computational data lead to these conclusions. The experimental observation that both re- actions were partially or fully inhibited by the presence of the radical inhibitor TEMPO argues against a concerted pathway for oxidative ad- dition of BrCH2CN (2-Br) to ionic Cu(I) com- plex [Ph4P]+[Cu(CF3)2]− (1a) or for oxidative addition of ClCH2CN (2-Cl) to neutral Cu(I) complex [(bpy)Cu(CF3)] (1b) (Fig. 4, pathway A). In addition, DFT calculations showed that the computed barrier (42.5 kcal/mol) for oxi- dative addition of BrCH2CN (2-Br) to complex 1a through a concerted pathway (pathway A) is much greater than that of the SN2-type path- way (pathway B; 22.5 kcal/mol) (Fig. 4) or XAT process (pathway D; 23.8 kcal/mol) (Fig. 4). Likewise, the computed barrier (36.3 kcal/mol) for oxidative addition of ClCH2CN (2-Cl) to complex 1b by pathway A is about 16.0 kcal/ mol greater than the lowest-energy alternative pathway, in this case the XAT process (path- way D; 20.3 kcal/mol). The experimental observation that the addi- tion of SET inhibitor 1,4-dinitrobenzene did not greatly affect either oxidative addition process argues against an OSET pathway (Fig. 4, pathway C). Moreover, the computed bar- riers for both reactions through an OSET pathway (39.3 kcal/mol and 24.3 kcal/mol, re- spectively) are 16.8 kcal/mol and 4.0 kcal/mol greater than the SN2-type pathway A for reac- tion with complex 1a or XAT pathway D for reaction with complex 1b. Thus, both the ex- perimental and computational data are incon- sistent with reaction through an OSET pathway (Fig. 4, pathway C). By contrast, the relative reactivity of the alkyl electrophiles toward the ionic Cu(I) complex 1a of ICH2CN > BrCH2CN >> TsOCH2CN >> ClCH2CN is consistent with an SN2-type path- way B. In addition, the reaction of the second- ary alkyl bromide 2-bromopropionitrile 2-Br-Me with ionic Cu(I) complex 1a, which is 1.5 times as slow as that of the less-hindered bromoace- tonitrile 2-Br [(1.15 ± 0.01) × 10−3 M−1·s−1 versus (1.78 ± 0.03) × 10−3 M−1· s−1], is inconsistent with OSET or XAT pathways (Fig. 4, pathways C and D). The reaction of a secondary alkyl halide through an OSET or XAT pathway would be faster than that of the primary alkyl halide because of the greater stability of the second- ary alkyl radical. Also consistent with reaction through an SN2 pathway B, the reactions of complex 1a Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E with BrCH2CN (2-Br) in polar solvents were faster than those in less-polar solvents, and the more sterically hindered 2-bromopropionitrile 2-Br-Me reacted more slowly than did 2-Br. The generation of TEMPO−CH2CN as a side product from the reaction of complex 1a with BrCH2CN (2-Br) in the presence of TEMPO suggests that an alternative but minor path- way also occurs during the reaction of the cuprate 1a. Consistent with this experimental observation, the barrier for a reaction of 1a with BrCH2CN (2-Br) through a XAT pathway D (23.8 kcal/mol) computed with DFT was only 1.3 kcal/mol greater than that of the SN2- type pathway B. Furthermore, the calculated barrier (22.5 kcal/mol for SN2-type pathway B and 23.8 kcal/mol for XAT pathway D) is close to the corresponding experimentally ob- served activation energy (21.3 kcal/mol), which provides additional evidence to support the proposed mechanism for reaction of 1a with haloacetonitrile. Our data for reaction of the neutral copper complex 1b are more consistent with XAT pathway D as the dominant mechanism. The considerable inhibiting effect of the addition of TEMPO on the reaction of chloroacetoni- trile 2-Cl with [(bpy)Cu(CF3)] (1b) suggests that a free radical •CH2CN is generated. The barrier for an OSET pathway C computed with DFT (24.3 kcal/mol) is much higher than that of the XAT process (pathway D, 20.3 kcal/mol). In addition, the calculated Gibbs free energy (DG; 20.3 kcal/mol) is close to the barrier (18.6 kcal/mol) determined experimentally, implying that a XAT process (pathway D) is the most likely pathway involving an alkyl rad- ical for oxidative addition of alkyl halides to the neutral Cu(I) complex 1b. Our experimental and computational data are most consistent with reaction of the alkyl halide directly with [(bpy)Cu(CF3)] (1b). The zero-order dependence on ligand is inconsistent with reversible ligand dissociation followed by oxidative addition (Fig. 4, pathway E). Fur- thermore, the computed DG for dissociation of bipyridine from complex 1b (17.5 kcal/mol) is accessible, but the computed barrier for a XAT process (pathway D) from ClCH2CN to the resulting ligandless [CuCF3] is an addi- tional 22.1 kcal/mol. The combination of these energies is much greater than the computed DG‡ (20.3 kcal/mol) for direct XAT to 1b (path- way D). To gain more insight into the oxidative ad- dition of the haloacetonitrile to Cu(I) complex 1a and 1b, we conducted natural population analysis (NPA) on reactants 1a, 1b, 2-Br, and 2-Cl; transition states TS-a-SN2, TS-a-XAT and TS-b-XAT; and products [LnCuIII(CF3) − or bpy) (figs. (X)(CH2CN)] (where Ln is CF3 S33 and S35). These studies showed that a small amount of positive charge accumulates on the copper in the transition states (+0.26 for 1a versus +0.35 for [CuIII(CF3)2(Br)(CH2CN)]−) during the reaction of 1a with 2-Br (53, 54) and that the negative charge of the CF3 moiety decreases considerably (−1.26 for 1a versus −0.59 for [CuIII(CF3)2(Br)(CH2CN)]−). Even though the change in charge to copper from starting complex to the intermediate after oxi- dative addition is small, the sum of the nega- tive charge of the bromide and cyanomethyl ligands (−CH2CN) is large (−0.76 e−). The same trend was also observed for the reaction of 2-Cl with 1b. These changes in charge show that electron density flows from the complexes 1a or 1b overall to the haloacetonitriles during the reaction, thus demonstrating that reaction of haloacetonitrile to 1a or 1b can be considered an oxidative addition process. Previous proposals for the mechanism of copper-catalyzed cross-coupling often involve a Cu(I)/Cu(II) redox cycle, on the basis of the experimental evidence for the presence of free radicals, which was typically deduced from quenching experiments with radical scav- engers or from racemization or rearrange- ments of radical clock substrates. Our studies suggest that a free radical could be involved in the oxidative addition of an alkyl halide to Cu (I) to form a Cu(III) intermediate in these pathways through XAT. Oxidative addition of the C(sp3)−X bond to a Cu(I) species is often considered to be the rate-limiting step of copper-catalyzed cross-coupling reactions of alkyl electrophiles, but studies of this ele- mentary step alone are rare, largely because of the inherent instability of the copper in- termediate. Thus, the example of oxidative addition of a C(sp3)−X bond to Cu(I) in the current study may help develop more efficient copper-catalyzed cross-coupling reactions of alkyl electrophiles. RE FERENCES AND NOTES 1. I. Beletskaya, A. V. Cheprakov, Coord. Chem. Rev. 248, 2337–2364 (2004). 2. G. Evano, N. Blanchard, M. Toumi, Chem. Rev. 108, 3054–3131 (2008). 3. F. Monnier, M. Taillefer, Angew. Chem. Int. Ed. 48, 6954–6971 (2009). 4. S. Bhunia, G. G. Pawar, S. V. Kumar, Y. Jiang, D. Ma, Angew. Chem. Int. Ed. 56, 16136–16179 (2017). 5. L.-J. Cheng, N. P. Mankad, Chem. Soc. Rev. 49, 8036–8064 (2020). 6. H. Zhou, Z.-L. Li, Q.-S. Gu, X.-Y. Liu, ACS Catal. 11, 7978–7986 (2021). 7. X.-Y. Dong et al., Nat. Chem. 11, 1158–1166 (2019). 8. X. Mo, B. Chen, G. Zhang, Angew. Chem. Int. Ed. 59, 13998–14002 (2020). 9. F. L. Wang et al., Nat. Chem. 14, 949–957 (2022). 10. E. J. Corey, G. H. Posner, J. Am. Chem. Soc. 89, 3911–3912 (1967). 11. D. H. Burns, J. D. Miller, H.-K. Chan, M. O. Delaney, J. Am. Chem. Soc. 119, 2125–2133 (1997). 12. G.-Z. Wang et al., Org. Lett. 17, 3682–3685 (2015). 13. C.-T. Yang, Z.-Q. Zhang, Y.-C. Liu, L. Liu, Angew. Chem. Int. Ed. 50, 3904–3907 (2011). 14. X.-L. Su et al., Angew. Chem. Int. Ed. 60, 380–384 (2021). 15. C. Li, B. Chen, X. Ma, X. Mo, G. Zhang, Angew. Chem. Int. Ed. 60, 2130–2134 (2021). 16. Q. M. Kainz et al., Science 351, 681–684 (2016). 17. C. Chen, J. C. Peters, G. C. Fu, Nature 596, 250–256 (2021). 18. N. Yoshikai, E. Nakamura, Chem. Rev. 112, 2339–2372 (2012). 19. C. Sambiagio, S. P. Marsden, A. J. Blacker, P. C. McGowan, Chem. Soc. Rev. 43, 3525–3550 (2014). 20. S.-J. Li, Y. Lan, Chem. Commun. 56, 6609–6619 (2020). 21. J. M. Ahn, T. S. Ratani, K. I. Hannoun, G. C. Fu, J. C. Peters, J. Am. Chem. Soc. 139, 12716–12723 (2017). 22. H. Lee et al., J. Am. Chem. Soc. 144, 4114–4123 (2022). 23. A. J. Hickman, M. S. Sanford, Nature 484, 177–185 (2012). 24. A. Casitas, X. Ribas, Chem. Sci. 4, 2301–2318 (2013). 25. L. Liu, Z. Xi, Chin. J. Chem. 36, 1213–1221 (2018). 26. H. Liu, Q. Shen, Coord. Chem. Rev. 442, 213923 (2021). 27. S. H. Bertz, S. Cope, M. Murphy, C. A. Ogle, B. J. Taylor, J. Am. Chem. Soc. 129, 7208–7209 (2007). 28. S. H. Bertz, S. Cope, D. Dorton, M. Murphy, C. A. Ogle, Angew. Chem. Int. Ed. 46, 7082–7085 (2007). 29. E. R. Bartholomew et al., Chem. Commun. (10): 1176–1177 (2008). 30. T. Gärtner, W. Henze, R. M. Gschwind, J. Am. Chem. Soc. 129, 11362–11363 (2007). 31. P. J. Brothers, W. R. Roper, Chem. Rev. 88, 1293–1326 (1988). 32. H. Shen et al., J. Am. Chem. Soc. 139, 9843–9846 (2017). 33. M. Paeth et al., J. Am. Chem. Soc. 141, 3153–3159 (2019). 34. S. Liu et al., J. Am. Chem. Soc. 142, 9785–9791 (2020). 35. A. Zanardi, M. A. Novikov, E. Martin, J. Benet-Buchholz, V. V. Grushin, J. Am. Chem. Soc. 133, 20901–20913 (2011). 36. H. Morimoto, T. Tsubogo, N. D. Litvinas, J. F. Hartwig, Angew. Chem. Int. Ed. 50, 3793–3798 (2011). 37. G. G. Dubinina, H. Furutachi, D. A. Vicic, J. Am. Chem. Soc. 130, 8600–8601 (2008). 38. H. Liu, Q. Shen, Org. Chem. Front. 6, 2324–2328 (2019). 39. S. Martínez de Salinas et al., Chem. Eur. J. 25, 9390–9394 (2019). 40. J. W. Tye, Z. Weng, A. M. Johns, C. D. Incarvito, J. F. Hartwig, J. Am. Chem. Soc. 130, 9971–9983 (2008). 41. J. W. Tye, Z. Weng, R. Giri, J. F. Hartwig, Angew. Chem. Int. Ed. 49, 2185–2189 (2010). 42. A. A. Isse, A. Gennaro, J. Phys. Chem. A 108, 4180–4186 (2004). 43. L. M. Rendina, R. J. Puddephatt, Chem. Rev. 97, 1735–1754 (1997). 44. J. K. Stille, K. S. Y. Lau, J. Am. Chem. Soc. 98, 5841–5849 (1976). 45. G. M. Whitesides, W. F. Fischer Jr., J. San Filippo Jr., R. W. Bashe, H. O. House, J. Am. Chem. Soc. 91, 4871–4882 (1969). 46. R. G. Pearson, C. D. Gregory, J. Am. Chem. Soc. 98, 4098–4104 (1976). 47. J. D. Sears, P. G. N. Neate, M. L. Neidig, J. Am. Chem. Soc. 140, 11872–11883 (2018). 48. G. D. Jones et al., J. Am. Chem. Soc. 128, 13175–13183 (2006). 49. J. B. Diccianni, J. Katigbak, C. Hu, T. Diao, J. Am. Chem. Soc. 141, 1788–1796 (2019). 50. C. Fang et al., J. Am. Chem. Soc. 141, 7486–7497 (2019). 51. K. Matyjaszewski, J. Xia, Chem. Rev. 101, 2921–2990 (2001). 52. P. Novák, A. Lishchynskyi, V. V. Grushin, J. Am. Chem. Soc. 134, 16167–16170 (2012). 53. I. M. DiMucci et al., J. Am. Chem. Soc. 141, 18508–18520 (2019). 54. J. Shearer, D. Vasiliauskas, K. M. Lancaster, Chem. Commun. 59, 98–101 (2022). AC KNOWLED GME NTS Funding: Q.S. gratefully acknowledges financial support from the National Key Research and Development Program of China (2021YFF0701700) and the National Natural Science Foundation of China (22061160465). X-S.X. acknowledges financial support from the National Natural Science Foundation of China (22122104). J.F.H. acknowledges funding from the NIH (R35GM130387). Y. Luo thanks Syngenta for a PhD scholarship. Author contributions: Y. Luo performed the experiments and analyzed experimental data. Y. Li performed the DFT calculations. J.W. assisted in the kinetic studies. X.-S.X., J.F.H., and Q.S. directed the research. Y. Luo, J.F.H., and Q.S. wrote the manuscript. Competing interests: The authors declare that they have no competing interests. Data materials and availability: Crystallographic data are available free Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E of charge from the Cambridge Crystallographic Data Centre (CCDC) under CCDC 2236874 (3a), 2236875 (3b), and 2236887 (trans-4). All other data reported in this paper are available in the manuscript or the supplementary materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg9232 Materials and Methods Figs. S1 to S36 Tables S1 to S15 References (55–76) Submitted 26 May 2023; accepted 10 August 2023 10.1126/science.adg9232 Luo et al., Science 381, 1072–1079 (2023) 8 September 2023 8 of 8
10.1126_science.adg9091
RES EARCH R E S E A R C H A R T I C L E S U M M A R Y ◥ ANTIMICROBIAL PROTEINS The mechanism of the phage-encoded protein antibiotic from FX174 Anna K. Orta†, Nadia Riera†, Yancheng E. Li, Shiho Tanaka, Hyun Gi Yun, Lada Klaic, William M. Clemons Jr.* INTRODUCTION: As part of their life cycle, viruses must escape from their hosts. For bacteriophages, this process requires breaching the bacterial cell wall and the peptidoglycan layer. Small, single- strand DNA and RNA phages have evolved single- gene lysis proteins that disrupt peptidoglycan biosynthesis and trigger cell lysis, a mechanism also used by some antibiotics. The best-studied of these is protein E from the historically impor- tant phage ΦX174, a 91–amino acid protein that contains a conserved N-terminal transmembrane helix and an extended cytoplasmic C-terminus. RATIONALE: Mutagenesis studies have dem- onstrated that the protein E transmembrane domain is sufficient for lysis, with functional mutants mapping to one face of the single transmembrane helix. The strongest pheno- types were observed for conserved prolines in this domain of protein E, with one proline being essential for lysis. Wild-type protein E requires the constitutively expressed bacterial chaper- one SlyD (sensitivity to lysis D), which contains a prolyl isomerase domain and a chaperone domain. Previous work had identified mutants The escape of FX174 from its bacterial host. In the membrane is the YES complex [E. coli enzyme MraY (cyan), phage protein E (yellow), and E. coli chaperone SlyD (purple)] where protein E disrupts peptidoglycan synthesis by inhibiting MraY and allowing breaching of the cell wall (tan). Assembled FX174 phage particles are shown in gray. In the background are bacterial cells that are lysing at their septal division point. in MraY that confer resistance to protein E– mediated lysis. MraY catalyzes the synthesis of a peptidoglycan precursor from a soluble nucleoside-sugar-peptide and a phospholipid substrate. Although many functional ques- tions remain about MraY, here we sought to understand how a phage protein could spe- cifically inhibit this key enzyme and what role SlyD performed in the process. RESULTS: We established that phage protein E forms a stable complex with the Escherichia coli proteins MraY and SlyD, designated the YES complex (MraY, protein E, SlyD). After solubilizing in detergent, we determined struc- tures of the YES complex by single-particle elec- tron cryo–microscopy. In our structures, the MraY dimer adopts a back-to-back orientation with the membrane-exposed active sites facing away from each other. Two SlyD molecules can be fit into the density on the cytoplasmic side. The two MraY and SlyD chains are bridged by two protein E molecules. The transmembrane helix of protein E occupies a groove on MraY corresponding to the putative binding site of the lipid substrate. At the cytoplasmic surface, the transmembrane domain of protein E bends to cross the active site, followed by an a-helix that extends to the chaperone domain of SlyD. The C- terminus of protein E continues through a peptide-binding groove in the second SlyD and into the prolyl isomerase active site. The conse- quence is that protein E inhibits MraY by block- ing lipid access to the active site. Previous protein E mutant phenotypes are explained by the struc- ture including the essential proline that allows for a kink in the transmembrane helix. For MraY, we can resolve the full chain including the conserved loop between TM1 and TM2. We show that the N terminus of MraY forms an a-helical stack with TM2 and identify ordered lipids on the protein surface. Finally, we provide evidence that the role of SlyD is to stabilize the complex. CONCLUSION: Protein E directly inhibits MraY by obstructing the MraY active site in a stable complex with SlyD. This structure resolves key questions about how the model phage FX174 kills bacteria and escapes the cell. In the FX174 genome, the gene encoding for protein E is evolutionarily constrained by gene D, in which it is embedded. The YES complex provides a route for the rational design of protein E beyond this gene D restraint. Single-gene lysis proteins, like protein E, serve as useful models for the development of antibacterial therapies.▪ The list of author affiliations is available in the full article. *Corresponding author. Email: clemons@caltech.edu †These authors contributed equally to this work. Cite this article as A. K. Orta et al., Science 381, eadg9091 (2023). DOI: 10.1126/science.adg9091 READ THE FULL ARTICLE AT https://doi.org/10.1126/science.adg9091 Orta et al., Science 381, 180 (2023) 14 July 2023 1 of 1 RES EARCH R E S E A R C H A R T I C L E ◥ ANTIMICROBIAL PROTEINS The mechanism of the phage-encoded protein antibiotic from FX174 Anna K. Orta1†, Nadia Riera1†, Yancheng E. Li1, Shiho Tanaka1, Hyun Gi Yun1, Lada Klaic1, William M. Clemons Jr.1* The historically important phage FX174 kills its host bacteria by encoding a 91-residue protein antibiotic called protein E. Using single-particle electron cryo–microscopy, we demonstrate that protein E bridges two bacterial proteins to form the transmembrane YES complex [MraY, protein E, sensitivity to lysis D (SlyD)]. Protein E inhibits peptidoglycan biosynthesis by obstructing the MraY active site leading to loss of lipid I production. We experimentally validate this result for two different viral species, providing a clear model for bacterial lysis and unifying previous experimental data. Additionally, we characterize the Escherichia coli MraY structure—revealing features of this essential enzyme—and the structure of the chaperone SlyD bound to a protein. Our structures provide insights into the mechanism of phage-mediated lysis and for structure-based design of phage therapeutics. T he full realization of phage therapy as a solution to the antimicrobial resist- ance problem requires a fundamental understanding of the mechanisms vi- ruses use to kill their host (1–4). Nearly 100 years ago, the first medical application of phages to treat infections used a cocktail containing the historically important phage FX174 (5). A rich source of critical discov- eries in molecular biology (6–9), FX174 is a member of the Bullavirinae subfamily within Microviridae. It is found broadly in environ- ments that contain coliform bacteria, such as the human gut (10), with Escherichia coli as its primary host. The 1977 publication of the FX174 single-stranded 5.4-kilobase DNA ge- nome was a milestone for genomics, reveal- ing 11 open reading frames (ORFs) (11). Most notable was the gene for lysis, E, that is em- bedded within the ORF of the scaffolding gene D (12). Expression of E alone is sufficient to kill bacteria (13). Although a variety of lysis mech- anisms have been proposed (14–16), the most likely is that the product of E, protein E, disrupts peptidoglycan (PG) biosynthesis, culminating in a breach in the cell wall. In particular, pro- tein E was demonstrated to directly inhibit the integral membrane enzyme MraY (16) depen- dent on the cytoplasmic chaperone SlyD (sen- sitivity to lysis D) (17, 18). Phage-derived single-gene lysis (SGL) pro- teins, such as protein E, trigger cell lysis by inhibiting PG biosynthesis [reviewed in (19)] providing a route for killing bacteria. Protein E includes a conserved N-terminal transmem- brane domain (TMD), a cytoplasmic C-terminal 1Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, CA 91125, USA. *Corresponding author. Email: clemons@caltech.edu †These authors contributed equally to this work. domain with conserved positively charged residues in a predicted amphipathic helix, and an unstructured tail (Fig. 1A and fig. S1). Fusions of globular proteins to the C termi- nus of the TMD have demonstrated that the TMD is sufficient for lysis (17, 20, 21). Exten- sive mutational analysis identified key res- idues in protein E including the essential proline 21 (P21) (17, 20–23). The mechanism through which SlyD and MraY work with protein E to facilitate rupture of the cell wall remains unknown. MraY, an important target for antimicrobial drug discovery [reviewed in (24)], catalyzes the transfer of a phospho- MurNAc-pentapeptide from “Park’s nucleotide” (UDP-MurNAc-pentapeptide) onto the phos- phate group of the 55 carbon polyisoprenyl- phosphate (C55P) generating lipid I. X-ray crystal structures of MraY in detergent from the thermophilic Gram-negative Aquifex aeolicus (AaMraY) (25, 26) or the Gram-positive Enterocloster bolteae (EbMraY) (27) revealed a conserved homodimeric structure, with ten TMDs per subunit. The structures localized the active site in a vestibule on the cytoplasmic side of a membrane-exposed cleft formed by TMDs 4, 5, and 9, which is the predicted access site for the lipid substrate (28). The constitutively expressed but nonessential (29) metallocha- perone (30) SlyD contains two conserved core domains: an FK506-binding protein (FKBP) prolyl-isomerase domain and an insertion-in- flap (IF) domain that binds unfolded peptides by b-augmentation (18, 31–33). The disordered SlyD C terminus (31), although important for metal binding, can be deleted without affect- ing chaperone activity (34). In the present study, we resolve the mech- anism of peptidoglycan biosynthesis inhibition by protein E. We determined the structure of the dimeric heterotrimer YES complex (EcMraY, viral protein E, EcSlyD) revealing that protein E physically blocks the lipid substrate from accessing the MraY active site. This work provides mechanistic insight into all three proteins and suggests a path toward devel- opment of antibacterial agents. The structure of the YES complex We used the protein E sequence from either the original phage FX174 or the shorter protein E isoform from phage ID21 (91 and 76 residues respectively), both from within Bullavirinae (Fig. 1A and fig. S1). We coexpressed an affinity- tagged protein EFX174 and wild-type (WT) EcMraY. After purification, a stable complex formed a single peak during size-exclusion chromatography (SEC) but resolved into four bands on a gel, two of which corresponded to the endogenous EcSlyD (fig. S2A). A SlyD variant with the disordered C terminus removed res- cued lysis activity in an E. coli DslyD strain (fig. S2, B to D) and ran as a single band on a gel (Fig. 1B) (18, 30). This truncation, SlyD154, was used for subsequent work except where noted. The two protein E isoforms induced lysis at similar efficiencies when expressed in a WT E. coli strain (Fig. 1C). For purification, all three genes in the YES complex (WT EcMraY, protein E, and truncated EcSlyD) were recombinantly expressed together in the DslyD strain with a C-terminal affinity tag on protein E. The complex was extracted in detergent and ran as a single peak by SEC with all three proteins in an apparent stoichiometric complex (Fig. 1B). Structures for both the YESFX174 and YESID21 complexes were solved using single- particle electron cryo–microscopy (cryo-EM) (Fig. 1D and figs. S3 and S4). The final density maps were obtained following several rounds of data processing with heterogeneous and ho- mogeneous refinements (figs. S3 and S4). The final overall masked resolution was 3.5 Å for the YESID21 complex and 3.6 Å for the YESFX174 complex (fig. S5). Statistics are provided in table S2. For both structures, the resolution was higher for regions in and adjacent to the membrane (figs. S3 and S4). We unambiguously built 90% of the protein residues in the complex. Sequence differences between the protein E variants are visible in the density (fig. S5), gen- erally exposed on the surface of the complex. The YESID21 complex is used as the reference structure, except where noted. Within the density map, we could clearly distinguish two copies of each member of the YES complex (six separate proteins). When contoured to remove the detergent micelle, densities for 22 TMDs are clear, 20 of which are accounted for by the EcMraY dimer (Fig. 1, D and E). Most of the cytoplasmic density can be accounted for by two SlyD molecules, which is the most flexible region of the com- plex (fig. S6D). The remaining protein den- sity corresponds to two protein E molecules Orta et al., Science 381, eadg9091 (2023) 14 July 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E A Transmembrane Domain Amphipathic Helix B C 10 20 30 40 50 X174 ID21 MGHWT L S G I L A F L L L L S L L L P S L L I MF I P L T FRRP A S SWK A RS L QK I L L MA S S ME RWT L L D I L A F L L L L S L L L P S L L I MF I P S MY K QHA S LWK A RS L A K T L S MA S S MV RWT LWD T L A F L L L L S L L L P S L L I MF I P S T FK RP V S SWK A L N L RK T L L MA S S G4 ME HWT L S G I L A F L L L L S L F L P S L L I MF I P L TS K P P V S SWK V L S L P K TS S MV L N ID8 ME HWT L S G I L A F L L L L S L L L P S L L I MF I P L TS K P P V S SWK V L S L P K TS S MV L N * * * * * * * ID21 X174 G4 ID8 D 60 70 80 90 100 V R L K P L S S S R I P CV L RP DS K RR F A R L TP L S S S R TP CV L K QDS K K L V R L K P L NCS R L P CV Y A QE T L T F L L TQK K TCV K NY V QK E RCNV A P L K P L NCS P S L F L FA P E TK I L S V T L K Q TS V NS Y A L K V P CKG L A Q L RP L NCS P S L FS FV P A TK T L S V MP RQ T FV NNY V L K V S CK Y Y E K I NS P L WW * * MraYB MraYA Protein EA Micelle Lipids Empty ID21 X174 0.10 0.08 0.06 0.04 0.02 ) U A ( e c n a b r o s b A 0.00 0 20 40 60 Time (min) MraY SlyD154 EID21 MW 25- 20- 15- 10- E MraYB MraYA 4 7 2 3 1 9a 6 8 5 4 10 9b Protein EA Periplasm N Cytoplasm C SlyDB Protein EB SlyDA SlyDB Protein EB SlyDA Fig. 1. The structure of the YES complex. (A) An alignment of a representative subset of protein E isoforms. Residue coloring is based on ClustalW. Secondary structure elements are shown above the sequence. Sequences are ordered as in fig. S1. Residues highlighted with a (*) at the bottom of the alignment are discussed in the text. (B) SDS-PAGE gel of the purified YESID21 complex. (C) A lysis assay for protein E expression. Cells containing either empty vector (black line) or the protein E genes for ID21 (pink line) or FX174 (green line) were induced at time 0 and the absorbance at 600 nm was monitored over time. Error bars represent the standard deviation derived from n = 3. (D) Overview density maps of the YESID21 complex viewed in the plane of the membrane. The detergent micelle is highlighted by gray coloration. The higher contoured map shows the six components of the twofold complex with density for E. coli MraY (cyan), protein E (yellow), and E. coli SlyD (purple) highlighted. The pairs of each protein are distinguished as the B-subunit is light purple. The general coloring scheme is maintained throughout the figures. Density that is likely lipid is shown in transparent orange. (E) Illustrated representation of the YESID21 complex oriented and colored as in (D). Foreground TMDs for the MraYs are numbered and the bilayer is represented by lines. N terminus and C terminus of protein EA are labeled. that each contain a kinked TMD and a solu- ble domain that bridges between MraY and SlyD. Densities for lipids are visible around the membrane-exposed surface of the MraY dimer (Fig. 1D). For protein E, starting at the N terminus in the periplasm, the TMD binds in the groove formed between TM5 and TM9 of MraY end- ing on the cytoplasmic side where it makes a sharp turn into the active site pocket (Figs. 1E and 2A). The TMD is followed by an amphi- pathic helix that crosses the active site, parallel to the membrane, presenting a positive face toward MraY and a hydrophobic face toward a SlyD IF-domain. The C-terminal residues adopt an extended conformation that pri- marily interacts with the second SlyD. This results in a crossover point between the two protein Es with each contacting both SlyDs. Overall, the dimeric complex (two of each of the heterotrimers) has a near twofold sym- metry perpendicular to the plane of the mem- brane. The membrane and periplasmic facing regions overlay perfectly but the symmetry is broken at the cytoplasmic face where the C-termini of the two protein E molecules cross each other at different residues and the SlyDs adopt slightly different orientations (Fig. 1E and fig. S6). Following toward the end of protein E, we can see continuous backbone density that positions proline 65 in the ac- tive site of the FKBP domain. Beyond that, the density is insufficient to resolve the sequence and we see little difference between FX174 and ID21 (fig. S5). The interaction of protein E with MraY Functional studies have consistently revealed the requirement for a proline at position 21 (21, 22, 35). Our structure allows an elegant explanation for this requirement. The pro- tein E TMD binds in the cleft of MraY that is defined by the angled TM9b (Fig. 2A). The proline at position 21 breaks the hydrogen bonding and introduces a kink that allows bending of the TMD around TM9b following the groove to the active site. Mutation of P21 would result in loss of the kink favoring a straight TMD that could not bind in the groove. Residue P29, also completely conserved (Fig. 1A and fig. S1), creates a second bend that com- pletes the wrap around TM9b and a mutation at this position results in delayed lysis onset (21, 22). Additional alanine mutations in the TMD identified residues that result in de- layed lysis onset, postulated to decrease the binding affinity of protein E to MraY (21). In the present structure, most of these residues (L19, L20, L23, and M26) (Fig. 2C) make direct contact with MraY. The exception, F27, appears Orta et al., Science 381, eadg9091 (2023) 14 July 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. The interaction of protein E with EcMraY. (A) Illustration of the YESID21 complex as in Fig. 1E with a 90° rotation except that MraYA is color- ramped using the Viridis palette that goes from purple to yellow for the N- to C-terminus, respectively. Prolines are shown in red. The bilayer is modeled in white. The B-subunits and SlyDs are faded. Boxes indicate regions highlighted in panels (B to D). Side chains for protein E residues that contact MraY are shown as sticks. (B) as in (A) with positions of protein E resistance mutants in MraY highlighted (dark cyan). (C) as in (A) with residues at the interface highlighted as sticks. (D) The region where the two protein E molecules cross highlighting the asymmetry colored as in (A) with interacting residues shown as sticks. Density for the amphipathic helix of protein E shown as a blue mesh. SlyDs are removed for clarity. (E) Lysis assay of expressed protein EFX174 variants. Error bars represent the standard deviation derived from n = 3. (F) is similar to (D) for protein EFX174. to sterically position M26 into a tight inter- action with Y134 in MraY, which is conserved in most Gram-negative bacteria (fig. S7). At the cytoplasmic interface, protein E res- idues A36 through M50 form the amphipathic helix spanning the width of an MraY subunit. The hydrophilic face of this helix orients toward the membrane in the MraY active site. The helix contains conserved positively charged residues that interact with conserved negatively charged residues in MraY (Fig. 2D). An example is the K46 salt bridge where, in our lysis assay, a K46A mutation resulted in delayed lysis onset (Fig. 2E). Here, isoform differences can be visualized. For example, position 42 is a leucine in FX174 whereas in ID21 it is an arginine that forms a salt bridge with E335 in MraY (Fig. 1). The helix ends at the crossover between the two protein Es at residue V54 in protein EA and A51 in protein EB resulting in differing interactions with residues in MraY, such as the essential H326 (Fig. 2D). The Epos (plaques on DslyD) mutations, R3H and L19F, allow phage propagation in a DslyD background (23). The R3H mutant of protein E in FX174 results in a silent mutation in pro- tein D and is found native to other species such as ID21 (fig. S8). Previous work reported that this variant resulted in higher levels of protein E in the membrane (23). In the present structure (fig. S5, G and H), this residue does not make specific contact with MraY. It is likely that the loss of a positive charge in the periplasm favors a higher percentage of pro- tein E molecules that are correctly inserted in the membrane as a result of the positive-inside rule. L19F does not result in higher levels of protein E (21) and likely hastens lysis onset by having a higher affinity to MraY. Another phenylalanine mutation at the interface, L23F, also hastens lysis onset likely by higher affi- nity, although this is not a general rule as other leucine to phenylalanine mutants did not af- fect lysis (21). Both L19 and L23 are near the conserved F182 in MraY and may add addi- tional stability through aromatic p interactions (Fig. 2C). The opposite mutation, phenylala- nine to leucine, can show loss of binding. For example, the F288L mutation in MraY (16) is located at the interface with protein E and re- sults in a loss of lysis, most likely due to lower affinity. All the mutations in MraY that allow re- sistance to protein E–mediated lysis (P170L, DL172, G186S, F288L, V291M) (16, 35) are at the interface with protein E (Fig. 2B). The mutant F288L, as noted above, lowers affi- nity to protein E. Residue G186 is at the near- est approach between the two proteins and a mutation to serine would prevent protein E binding. The mutant V291M lies directly at the interface near L19 in protein E, although a specific effect for this mutation is not clear. Finally, P170L and DL172 are located within periplasmic loop 4-5, which interacts with the conserved N terminus of protein E. Although more resistant to lysis, these two mutants are predicted to still bind protein E, albeit with lower affinity (35). The mechanism of inhibition of MraY by protein E The YES complex structure allows us to pro- pose a simple mechanism for inhibition of MraY by protein E. Superposition with previ- ous MraY inhibitor complexes (26, 27) on the YES complex (Fig. 3) shows that the predicted path of the polyisoprenyl chain of C55P is the groove formed between TM5 and TM9, which is occluded by protein E. Therefore, one mech- anism of inhibition is that protein E prevents access of the lipid substrate to the active site (Fig. 3A). Protein E is a noncompetitive inhib- itor of Park’s nucleotide (36) and, accord- ingly, the pocket that binds the nucleoside is fully accessible in the present structure (Fig. 3B). Loop 9-10 in MraY contains catalytic his- tidines that must move toward the binding pocket to facilitate catalysis by completing the active site (28). The cytoplasmic helix of protein E separates loop 9-10 from the rest of the active site, blocking this transition and providing a second mechanism of inhibition. Overall, protein E blocks access of the lipid substrate and prevents formation of the active site upon substrate binding. Orta et al., Science 381, eadg9091 (2023) 14 July 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E A carbacaprazamycin tunicamycin A EcMraY B 5 Protein E 4 AaMraY AaMraY (inhibited) Loop 9-10 9b B lipid binding site Loop 1-2 C RMSD - EcMraY 0.966 AaMraY(6OYH) 1.081 AaMraY(4J72) 1.024 EbMraY(5JNQ) * N-terminus Loop 3-4 * EcMraY Protein EA * Periplasm EbMraY * AaMraY 2 1 2 1 Cytoplasm nucleoside binding site Fig. 3. The mechanism of inhibition by protein E. (A) Accessible surface of MraY (colored as in Fig. 2A) viewed from the cytoplasm looking toward the active site cleft. Protein E is shown as an illustration. The structures for the inhibitors tunicamycin (maroon) and carbacaprazamycin (pink) are shown as sticks based on a structural alignment to EcMraYA with their respective complex structures [PDBID: 6OYH (26) and 5JNQ (27)]. (B) Similar to (A) from a slightly different angle highlighting the catalytic pocket. MraY (dark cyan) is shown as an illustration. The two substrate binding sites are highlighted by dashed boxes. Predicted catalytic residues in MraY are shown as sticks. Key structural features of MraY Although there are prior crystallographic MraY structures (25–27), our YES complex structure reveals E. coli MraY in a distinct structural context (fig. S9). There is agreement with the prior structures when comparing MraY mono- mers with backbone RMSDs around 1 Å (Fig. 4A and fig. S10). We can model previously disor- dered regions of MraY including all the cyto- plasmic loops that enclose the active site. Loop 1-2 likely adopts distinct conformations during the catalytic cycle and indeed has two slightly different conformations in our dimer (fig. S6C). Loop 9-10 adopts a conformation that is simi- lar to the reported small-molecule inhibitor co-crystal structures of MraY (26, 27, 37) but distinct from the apoprotein structure (Fig. 4 and fig. S10). This conformation of loop 9-10 Fig. 4. Features observed in the EM structure of E. coli MraY. (A) Cytoplasmic view of a structural alignment of EcMraY (dark cyan) against uninhibited AaMraY [green, PDBID:4J72 (25)] and carbacaprazamycin inhibited AaMraY [pink, PDBID:6OYH (26)]. RMSDs to monomer A of EcMraY are shown. The color scheme for the various MraY crystal structures is used throughout the figures. (B) A view in the plane of the membrane showing the region that includes TM1 and TM2 in backbone ribbons. (Left) The EcMraY structure colored from the N terminus through TM2 in the Viridis color scheme. (Right) The inhibited AaMraY structure from (A) and the EbMraY structure [gray, PDBID:5JNQ (27)]. Each structure is aligned to EcMraY. The location of each N terminus is indicated by an asterisk. (C) Accessible surface of EcMraY (cyan) and unmodeled densities (orange) that are likely lipids or detergent as well as a putative C55P molecule (purple, see below). The inset is a view of the periplasmic cavity viewed from a removed monomer. may be a general feature of MraY complexed with inhibitors. In prior crystal structures, the N terminus of MraY begins at either TM1 or is a helix that projects away from the structure in an orienta- tion incompatible with the bilayer (Fig. 4B). In the YES complex, the N-terminal end of the first helix (NTH) hydrogen-bonds to the C-terminal end of TM2, effectively forming a helical stack- ing structure (Fig. 4B and fig. S9C). A multiple sequence alignment across bacteria shows that the NTH in MraY is conserved across gram- negative bacteria but missing in Gram posi- tives (figs. S7 and S11). Although this feature is not found in the crystal structures, AlphaFold (38) predicts the NTH stacking for E. coli and other Gram-negative bacteria with slight dif- ferences in hydrogen bonding and orientation relative to the cryo-EM structure (fig. S11). For the AaMraY structures (25, 26), the positioning of the N terminus is likely a product of crys- tallization, as AlphaFold predicts the NTH stacks for this and the related Hydrogenivirga spe- cies (fig. S11). For Gram-positive bacteria, both the EbMraY and predicted structures lack the NTH (27) (Fig. 4B and fig. S11). Protein E as a general antibacterial protein Expression of protein E leads to lysis that results in a subset of killed bacteria becom- ing “ghosts”—essentially empty cell walls (15). Ghosts have been used in a variety of contexts including vaccine development [reviewed in (39)]. For ghosts to be made, protein E must inhibit the native MraY of the target bacteria. Although phage FX174 is restricted to E. coli, many bacteria have been probed for ghost for- mation by expression of protein E. All Gram- negative bacteria tested (fig. S11A) resulted in the formation of ghosts [reviewed in (40)]. By contrast, protein E was unable to cause lysis in the Gram-positive Staphylococcus carnosus (41). Expression of the Bacillus subtilis MraY in E. coli prevented cell lysis, suggesting that it has a low affinity for protein E (35). Compar- ing the residues in EcMraY that contact pro- tein E to those in Gram-positive species, there are no single residue differences that easily explain the inability to inhibit MraY from gram- positive species. There are a few residues that may play a role—such as P170 in EcMraY which confers resistance when mutated to leucine— that are absent in Gram-positive MraYs (fig. S7). Orta et al., Science 381, eadg9091 (2023) 14 July 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E EcMraY residue Y134, which forms a bridge with protein E M26 in our model (Fig. 2C), is also missing in Gram-positive species (fig. S7). Lipids bound to the YES complex The YES complex was solubilized from its na- tive environment and, for the YESID21 complex, the final detergent solution was supplemented with E. coli lipids. We observe lipid densities around the membrane surface of MraY (Fig. 1D and fig. S12, A and B). Prior work supports a functional role for anionic phospholipids with MraY including stabilizing the dimer (42–44). We observe substantial lipid density near the dimer interface with features consistent with glycerophospholipids, although not all density can be clearly assigned (fig. S12). As in previous structures (25, 27), there is a hydrophobic peri- plasmic cavity within the MraY dimer that con- tains unexplained density (Fig. 4C). Although this density has clear structure, we are unable to fit typical E. coli phospholipids. Recently, a study using native mass spec- trometry and molecular dynamics identified a binding site for C55P at the dimer interface with the phosphate headgroup predicted to form a salt bridge to R341 in EcMraY (44). In our structure, we observed a long tube of den- sity at the dimer interface (fig. S12A, purple) that ends at R341 and is consistent with C55P (fig. S12C). Although we cannot confidently iden- tify this lipid, the density would be consistent with C55P, although the role for this binding site remains to be determined. The role of SlyD in protein E–mediated lysis The amphipathic helix of protein E bridges the two E. coli proteins, MraY and SlyD, which make no specific contact with each other (Fig. 2A). The IF domain sits on the hydrophobic face of the protein E helix (Fig. 5A) and con- tacts the extended C terminus of the oppos- ing protein E, which then extends to bind the FKBP domain (Fig. 5B). This results in a bowtie-like interaction with each SlyD binding both protein E–soluble domains (Fig. 1E). There are no contacts between the two SlyD mol- ecules (fig. S6D). These protein E interfaces validate structural studies of SlyD where pep- tides bind to each of the interfaces seen here including b-augmentation in a groove in the IF domain (Fig. 5, C and D) (32). The IF domain binding to an a-helix may be an important chaperoning interaction. The FKBP domain is well-ordered but adopts a range of orientations in our particles (fig. S13) and consequently has the lowest resolution (fig. S3). This flexibility relative to the IF domain is consistent with the NMR structure of SlyD where the orientations of the two domains are not constrained by each other (31). The only contact of the FKBP do- main to the rest of the YES complex is the flexible linker to the IF domain and binding to the extended C terminus of protein E. We ob- serve continuous backbone density that allows us to place P65 at the FKBP active site. As seen before for some SlyD-bound peptides (32), pro- tein E binding to the FKBP domain adopts a noncanonical orientation. Protein E P65 is not A C B D Y68 P65 Y13 P58 F96 Y63 Pro Y13 S2 peptide F91 Pro TtSlyD PDBID: 7OXI Fig. 5. Interactions between protein E and SlyD. (A) The amphipathic helix of protein E (yellow) as a illustration with side chains contacting SlyD shown as sticks. SlyD (purple) is shown as an illustration along with a transparent accessible surface (B) Full view of SlyD bound to two protein E molecules shown in different shades of yellow. (C) As in (A) highlighting the b-augmentation of the extended C-terminal protein E. (D) Structures of SlyD from the YESID21 complex and TtSlyD [PDBID:7OXI (33)] (green) aligned to the IF domains. The two S2 peptides bound to the TtSlyD are shown in pink. completely conserved (fig. S1), although other species have a proline at residue 63 which could reach the active site with additional tilting of the FKBP domain. A proline near this posi- tion may help to localize the chaperone to the complex and facilitate assembly. The lack of contacts between SlyD and MraY suggest that the soluble domain of protein E alone could form a complex with SlyD. We coexpressed EcSlyD with N-terminal trunca- tions of protein E from either ID21 (residues 33 to 76) or the shortest isoform a3 (residues 33 to 75). Both form complexes that could be purified by an affinity tag on the soluble do- main (fig. S2E). These observations point to a high-affinity interaction between SlyD and the C-terminal domain of protein E. Protein E is unstable in the absence of SlyD (18) and is rapidly degraded (23). It has been speculated that prolyl-isomerization is central to the lysis mechanism (22), but this idea is not supported by available evidence. Nonproline mutations in protein E can rescue lysis in a DslyD background (18, 45), as well as completely replace the cytoplasmic domain of protein E with an unrelated globular protein (20, 21, 45). We also performed our lysis assay with sev- eral SlyD variants in the DslyD strain (fig. S2C). As before, protein EFX174 was unable to pro- mote lysis in the absence of SlyD, however lysis could be rescued by expression of either EcSlyD or EcSlyD154. Thermus thermophilus (Tt) SlyD has high structural homology to EcSlyD (Fig. 5D and fig. S2, B, F, and G) and also res- cued lysis in the DslyD strain (fig. S2C). We generated an EcSlyD Y68K mutant that sub- stantially reduces prolyl-isomerase activity (46) and this too could rescue lysis. Finally, we purified a complex of protein E using the Epos rescue mutants (R3H and L19F) (45) in the DslyD strain. This YE complex was very unstable and aggregated according to SEC (fig. S2H). These results support the conclusion that the primary role of SlyD in promoting lysis is not prolyl-isomerization, but to protect protein E and stabilize the YES complex. This does not obviate a role for proline binding in complex assembly. Mutation of P65, in the FKBP active site, creates a slower-lysis phenotype (22), which may indicate that binding of these residues by SlyD is key to assembling the YES complex. Protein E is evolutionarily constrained Protein E arrived late in the evolution of FX174 and was overprinted into a +1 reading frame in the ORF for gene D (47, 48). The structure supports that this embedding constrains the evolution of protein E (12). Considering the sequence changes across protein E isoforms, we note that from the N terminus through residue 70, with few exceptions, each position that is not completely conserved is either si- lent or that there is only a slight change in protein D (fig. S8A). The silent changes vary Orta et al., Science 381, eadg9091 (2023) 14 July 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E the codon’s second position, which is the wob- ble position in the overlapping codon for pro- tein D. For example, residue W7 in FX174 is replaced with a Ser or Leu in other species. Although seemingly substantial changes, each is coded by the sequence UXG and all three variations at this position are sampled. These variable positions generally do not contact MraY. An exception is found in the G4 isoform where a phenyalanine occurs at position 19, as in the Epos mutant, which results in a change from alanine to serine at position 79 in protein D. This places a polar residue in the hydro- phobic core of protein D (fig. S8). This single nucleotide change increases protein E affinity to MraY but likely lowers the stability of pro- tein D and the overall fitness of the virus. In- troduction of this mutation results in smaller plaque formation relative to the wild type (23). The C terminus of protein D and the exten- sions in the longer isoforms of protein E are likely disordered and less constrained, hence the increased sequence variability (fig. S8). Conclusions The mechanism for FX174 inhibition of pep- tidoglycan biosynthesis is suprisingly simple. Protein E binds in the active site cleft prevent- ing access of the lipid substrate and blocking conformational changes needed for catalysis (fig. S14). The YES complex structure provides a template for interpreting all the previous results for lysis by FX174. Notable among these are that the extensive functional mutations in protein E and MraY map to the interface between the two proteins. SlyD binds to the cytoplasmic domain of protein E to stabilize the MraY/protein E complex. As we demon- strated, the requirement for SlyD is based on the protein binding properties, as distantly re- lated SlyD homologs can rescue lysis in a DslyD strain. SlyD in this case does not serve directly in the mechanism of inhibition of MraY and can be functionally replaced. Our structures offer some insights for future work applying SGL proteins as therapies. Pro- tein E binding to EcMraY identifies a distinct mechanism from current inhibitors. Blocking of the lipid substrate access to the active site in MraY could be exploited by small molecules, with structure-based drug design guided by the protein E structure. Expanding protein E efficacy to Gram-positive bacteria and, potentially, other related polyisoprenyl-phosphotransferases such as bacterial WecA or the mammalian DPAGT1 could further increase applicability. SGL proteins, which have distinctly evolved many times (49), can also be used to improve the antibacterial potency of bacteriophages. The use of phages for medical therapies, although known for a hundred years, has attracted re- cent attention (1). Improving the efficiency of FX174 protein E based on structural infor- mation may help these therapies achieve their greatest potential in the fight against antimi- crobial-resistant pathogens. Materials and Methods Coexpression of EcMraY, protein E, and EcSlyD DslyD BL21(DE3)-competent cells were co- transformed with pET22b-SlyD154 and either pRSFDuetEcMraY-EID21 or pRSFDuet-EcMraY- EFX174 and plated in LB-agar containing 35 mg/ml Kanamycin and 100 mg/mL Ampicillin. Our pET22b-SlyD154 construct expresses E. coli SlyD, modified by the removal of the flexible C- terminus. The pRSFDuet-EcMraY-EID21 plas- mid contains the ID21 isoform of protein E, along with a wild-type EcMraY to prevent cell lysis from the overexpression of protein E. Cells were grown in 2xYT media at 37°C, 225 r.p.m., and induced at an OD600 of 0.9 with 0.4mM IPTG at 18°C overnight. The culture was har- vested by centrifugation for 10 minutes at 9000xg, 4°C then frozen or used immediately for purification. Purification of the YES complex The cells were resuspended in lysis buffer (20mM Tris-HCl pH 7.5, 300 mM NaCl, 10% Glycerol, 5mM b-mercaptoethanol (bME), 0.1mM PMSF, 0.1mM Benzamidine) and homogenized using a M-110L microfluidizer (Microfluidics). The lysate was cleared by a 20-minute centrifugation at a speed of 22,000xg. The supernatant was then centrifugated at 167,424xg and the resulting membrane pellet was then solubilized in the extraction buffer [10 mM HEPES pH 7.5, 300 mM NaCl, 5% Glycerol, 5mM bME, 0.1mM phenyl- methylsulfonyl fluoride (PMSF), 0.1mM ben- zamidine, 10 mM imidazole and 1% dodecyl 4-O- a-D-glucopyranosyl-b-D-glucopyranoside (DDM)] After allowing for extraction for 1.5 hours at 4°C, the solution was centrifuged at 167,424xg for 30 minutes and the remaining lysate was mixed with 1mL NiNTA resin (Qiagen, Alameda, CA) then nutated at 4°C for two hours. This solution was loaded onto a gravity column and then washed with five column volumes of wash buffer (10 mM HEPES pH 7.5, 150 mM NaCl, 5% glycerol, 5mM bME, & 0.03% DDM) with 10mM imidazole followed by five column vol- umes of wash buffer with 30 mM imidazole. The YES complex was eluted in 20mL of wash buffer containing 200 mM imidazole. The final purification step was SEC (Superdex 200 10/300 GL, Millipore Sigma) in 10mM HEPES pH 7.5, 75 mM NaCl, 5% Glycerol, 5mM bME and 0.03% DDM. Fractions were assessed by SDS-PAGE and directly used for cryo-EM sample preparation. Co-expression of EcMraY and protein E in various SlyD backgrounds The pRSFDuet-E-EcMraY and pRSFDuet-Epos- EcMraY expression vectors were transformed into BL21-Star cells (Novagen). Similarly, the pRSFDuet-E(C-term)-SlyD154 was transformed into SlyD-knockout cells. The cultures were grown at 37 °C to an OD600 0.8 and induced with 1 mM IPTG. Induced cultures were grown for 3 hours followed by harvesting by centrifu- gation at 9,000xg for 20 min. Cell pellets were resuspended in lysis buffer and lysed by son- ication. The lysate was then cleared by cen- trifugation at 22,000xg, followed by a second centrifugation at 234,78xg for 1 hour to isolate the membrane fraction. The complex was ex- tracted in 20 mM Tris-HCl pH 7.5, 300 mM NaCl, 10% Glycerol, 10 mM Imidazole, and 1% n-Decyl- b-Maltoside (DM) and incubated at 4°C for 1.5 hours. The debris was cleared by centrifu- gation at 234,788xg for 30 min. The sample was incubated with 1 mL NiNTA resin for 1 hour, followed by a wash with 50 column volumes lysis buffer with 30mM Imidazole. The protein E complexes were similarly eluted in 300mM Imidazole. The elutions were concentrated and further purified by size exclusion chromatogra- phy (Superdex 200 10/300 GL, Millipore Sigma). Lysis assays of WT protein E FX174 and ID21 LEMO DE3 competent cells were transformed with a pRSF-Duet vector either empty, with protein EFX174, or protein EID21. Cultures were grown to an OD600 of 0.2 and inoculated into a Corning 96-well Clear Flat Bottom plates in 100mL triplicate aliquots and induced as de- scribed previously. Cultures were incubated at 37C with orbital shaking at 220rpm using an Infinite M Nano+ (Tecan, Switzerland). Readings were taken in 5-minute intervals for 90 minutes. Lysis assay for protein E constructs LEMO DE3 competent cells (New England Biolabs, MA, USA) were transformed with a pRSFDuet vector either empty, with C-terminally FLAG tagged protein EFX174 variants (WT, P21A, K46A). The lysis assays were performed in three biological replicates as previously de- scribed (21). Absorbance readings were recorded in 5-minute intervals for 1 hour and 30 minutes. Manual readings were taken using a Biowave Cell Density Meter CO8000. The values were plotted using GraphPad Prism version 9.1.1 for macOS. Lysis assays based on SlyD variants DslyD (18) cells were transformed with either a control empty pRSF-Duet vector or pRSFDuet- ProteinFX174 and either pET22b-EcSlyD, pET22b- SlyD154, pET22b-EcSlyD Y68K, or pET22b-Thermus thermophilus SlyD. Cultures were grown in 2xYT media at 37°C and induced with 0.4mM IPTG once at an OD600 of 0.2. Absorbance measure- ments were manually recorded in 5-minute intervals for 70 minutes. Similarly, DslyD cells were transformed with either a control empty pRSF-Duet vector or pRSF-Duet-ProteinID21 either alone, with pET22bEcSlyD, or with pET22b- EcSlyD154 and induced with 0.4mM IPTG. Read- ings were recorded using an Infinite M Nano+ plate reader as described above. Orta et al., Science 381, eadg9091 (2023) 14 July 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E Sample preparation for cryoEM The YES complex was diluted to 5.0 mg/mL in 10 mM HEPES pH 7.5, 75 mM NaCl, 5% Glyc- erol, 5mM bME and 0.03% DDM. Additionally, the YESID21 sample was supplemented with 1mM E. coli total lipid extract (Avanti Polar Lipids, 100600P). Quantifoil holey carbon films R1.2/1.3 300 Mesh, Copper (Quantifoil, Micro Tools GmbH) grids were glow discharged with a 2-minute 20Å plasma current using a Pelco easiGlow, Emitech K100X. Grids were prepared using a Vitrobot (FEI Vitrobot Mark v4 x2, Mark v3) by applying 3mL of sample onto the grid followed by a 3.5 second blot using a +8-blot force and plunge frozen into liquid ethane. Data acquisition and analysis The grids were imaged in a 300 kV cryo-TEM microscope equipped with a Gatan K3 6k x 4k direct electron detector and a Gatan Energy Filter (slit width 20eV) in super-resolution mode using Serial EM. Datasets were collected at a 105k magnification with a pixel size of 0.416 Å/pixel. Movies with 40 frames were recorded with a total exposure dose of 60 e-/Å2 and a defocus range of -1.0 to -2.5 mm. For the YESID21 complex, a total of 12,070 movies were recorded. Movies were normalized by gain reference and mo- tion corrected using the patch motion correc- tion built in function in cryosparc (v3.3.2) with a twofold bin that resulted in a pixel size of 0.832 Å/pixel (50). The contrast transfer func- tion (CTF) was estimated using CTFFIND4 (51). Micrographs were manually curated, and low- quality images were removed for further analy- ses. A total of 2,462,335 particles were obtained followed by the generation of 6 ab-initio mod- els. Out of the 6 models, two models are se- lected for classification into “good” and “trash” volumes. All particles were then sorted in these two volumes through heterogeneous refinement using particles extracted with a 4x bin, which produced 6,589,696 good particles. Heteroge- neous refinement was used in an iterative man- ner to sort the particles into the 5 volumes (4 good and 1 trash). The 1,151,777 good particles were used for non-uniform homogeneous re- finement to generate a higher resolution vol- ume. The particles were then extracted with a 3x bin and sorted into 4 iterations of the higher resolution volume and 1 trash volume. Iterative rounds of heterogeneous refinement at 3x bin produced 935,754 particles. Particles were then extracted in a 2x bin and heteroge- neously refined into either high- or low-resolution volumes. At this point, discerning features in the soluble region of the model were used to select the most complete volumes. The volumes were individually refined through non-uniform refinement and the particles that composed the volumes with most complete and highest reso- lution were used. A total of 122,452 particles were used for the most complete model obtained upon non-uniform refinement. The FSC-masked resolution was 3.5Å, while the unmasked reso- lution was 3.9Å. The half-maps were then used for post-processing through DeepEMhancer (52) with the high-resolution model selected for our most complete density map. Post-processing through DeepEMhancer removed the micelle and improved the features on the soluble por- tions of the map, however the lipid densities were also removed. The lipid densities described in this work are those of the YESID21 map before post-processing. Notably, the dimer-interface lipid density was also present in the YESFX174 density map without the supplementation of E. coli lipid extract. Figure 1D uses the densities before post-processing for the MraY dimer, mi- celle and lipid densities, and the DeepEMhancer post-processed map for protein E and SlyD. Supplemental figures S9 and S12 were made with the map before DeepEMhancer sharpen- ing. The YESFX174 complex dataset was pro- cessed in this same manner. A total of 10,798 movies were recorded. The model for the YESID21 complex was then used as a template for tem- plate picking, from which 1,516,368 particles were picked and curated. Following gradual un-binning and sorting into good and trash volumes, 155,270 particles were used for the final iteration of non-uniform refinement. The local resolution of both maps was performed on cryosparc (v3.3.2). The half-maps were then post-processed using DeepEMhancer as de- scribed previously. Model building For starting models we used the Aquifex aeolicus un-bound structure (PDBID:4J72) (25) for EcMraY and the E. coli NMR structure (PDBID:2K8I) (31) for EcSlyD which were fit using phenix.dock. SlyD was then split into its two domains, IF and FKBP, at residues Y68 and G127. Protein E was modeled de novo up to residue P65 using Coot 0.8.9.2. The structures of the EcMraY, pro- tein E, and EcSlyD-IF domain were refined using phenix.real space refinement and ISOLDE 1.6, and validated with PHENIX-1.19.2. After the re- finements of EcMraY, protein E, and the EcSlyD-IF domain were completed, the FKBP domains were docked into density using ChimeraX. The complete YES complex structure was then re- fined using PHENIX-1.19.2 and ISOLDE 1.6. RMSDs were calculated using ChimeraX Match- maker chain alignment. Structure figures were made using ChimeraX and sequence alignments using Jalview and ClustalW (53–55). RE FERENCES AND NOTES 1. S. A. Strathdee, G. F. Hatfull, V. K. Mutalik, R. T. Schooley, Phage therapy: From biological mechanisms to future directions. Cell 186, 17–31 (2023). doi: 10.1016/ j.cell.2022.11.017; pmid: 36608652 2. R. T. Schooley et al., Development and Use of Personalized Bacteriophage-Based Therapeutic Cocktails To Treat a Patient with a Disseminated Resistant Acinetobacter baumannii Infection. Antimicrob. Agents Chemother. 61, e00954–e17 (2017). doi: 10.1128/AAC.00954-17; pmid: 28807909 3. C. Ferriol-González, P. Domingo-Calap, Phage Therapy in Livestock and Companion Animals. Antibiotics 10, 559 (2021). doi: 10.3390/antibiotics10050559; pmid: 34064754 4. K. E. Kortright, B. K. Chan, J. L. Koff, P. E. Turner, Phage therapy: A renewed approach to combat antibiotic-resistant bacteria. Cell Host Microbe 25, 219–232 (2019). doi: 10.1016/ j.chom.2019.01.014; pmid: 30763536 5. F. D. Herelle, [The bacteriophage]. Atomes 3, 399–403 (1948). pmid: 18104413 6. R. L. Sinsheimer, A single-stranded deoxyribonucleic acid from bacteriophage FX174. J. Mol. Biol. 1, 43–IN6 (1959). doi: 10.1016/S0022-2836(59)80006-1 7. C. A. Hutchison 3rd, R. L. Sinsheimer, The process of infection with bacteriophage phi-X174: X. Mutations in a phi-X Lysis gene. J. Mol. Biol. 18, 429–447 (1966). doi: 10.1016/ S0022-2836(66)80035-9; pmid: 5968177 8. R. M. Benbow, C. A. HutchisonIII, J. D. Fabricant, R. L. Sinsheimer, Genetic map of bacteriophage FX174. J. Virol. 7, 549–558 (1971). doi: 10.1128/jvi.7.5.549-558.1971; pmid: 16789129 9. H. O. Smith, C. A. Hutchison3rd, C. Pfannkoch, J. C. Venter, Generating a synthetic genome by whole genome assembly: phiX174 bacteriophage from synthetic oligonucleotides. Proc. Natl. Acad. Sci. U.S.A. 100, 15440–15445 (2003). doi: 10.1073/pnas.2237126100; pmid: 14657399 10. M.-S. Kim, E.-J. Park, S. W. Roh, J.-W. Bae, Diversity and abundance of single-stranded DNA viruses in human feces. Appl. Environ. Microbiol. 77, 8062–8070 (2011). doi: 10.1128/ AEM.06331-11; pmid: 21948823 11. F. Sanger et al., Nucleotide sequence of bacteriophage phi X174 DNA. Nature 265, 687–695 (1977). doi: 10.1038/ 265687a0; pmid: 870828 12. B. G. Barrell, G. M. Air, C. A. Hutchison 3rd, Overlapping genes in bacteriophage phiX174. Nature 264, 34–41 (1976). doi: 10.1038/264034a0; pmid: 1004533 13. K. D. Young, R. Young, Lytic action of cloned phi X174 gene E. J. Virol. 44, 993–1002 (1982). doi: 10.1128/jvi.44.3.993- 1002.1982; pmid: 6294347 14. W. Lubitz, G. Halfmann, R. Plapp, Lysis of Escherichia coli after infection with phiX174 depends on the regulation of the cellular autolytic system. J. Gen. Microbiol. 130, 1079–1087 (1984). doi: 10.1099/00221287-130-5-1079; pmid: 6236279 15. A. Witte, G. Wanner, M. Sulzner, W. Lubitz, Dynamics of PhiX174 protein E-mediated lysis of Escherichia coli. Arch. Microbiol. 157, 381–388 (1992). doi: 10.1007/ BF00248685; pmid: 1534215 16. T. G. Bernhardt, W. D. Roof, R. Young, Genetic evidence that the bacteriophage phi X174 lysis protein inhibits cell wall synthesis. Proc. Natl. Acad. Sci. U.S.A. 97, 4297–4302 (2000). doi: 10.1073/pnas.97.8.4297; pmid: 10760296 17. D. Maratea, K. Young, R. Young, Deletion and fusion analysis of the phage phi X174 lysis gene E. Gene 40, 39–46 (1985). doi: 10.1016/0378-1119(85)90022-8; pmid: 2936651 18. W. D. Roof, S. M. Horne, K. D. Young, R. Young, slyD, a host gene required for phi X174 lysis, is related to the FK506- binding protein family of peptidyl-prolyl cis-trans-isomerases. J. Biol. Chem. 269, 2902–2910 (1994). doi: 10.1016/ S0021-9258(17)42027-8; pmid: 8300625 19. K. R. Chamakura, R. Young, Single-gene lysis in the metagenomic era. Curr. Opin. Microbiol. 56, 109–117 (2020). doi: 10.1016/j.mib.2020.09.015; pmid: 33075663 20. K. J. Buckley, M. Hayashi, Lytic activity localized to membrane- spanning region of phi X174 E protein. Mol. Gen. Genet. 204, 120–125 (1986). doi: 10.1007/BF00330198; pmid: 3018438 21. S. Tanaka, W. M. Clemons Jr., Minimal requirements for inhibition of MraY by lysis protein E from bacteriophage FX174. Mol. Microbiol. 85, 975–985 (2012). doi: 10.1111/ j.1365-2958.2012.08153.x; pmid: 22742425 22. A. Witte, G. Schrot, P. Schön, W. Lubitz, Proline 21, a residue within the a-helical domain of phiX174 lysis protein E, is required for its function in Escherichia coli. Mol. Microbiol. 26, 337–346 (2003). doi: 10.1046/j.1365-2958.1997.5781941.x; pmid: 9383158 23. T. G. Bernhardt, W. D. Roof, R. Young, The Escherichia coli FKBP-type PPIase SlyD is required for the stabilization of the E lysis protein of bacteriophage ϕ X174. Mol. Microbiol. 45, 99–108 (2002). doi: 10.1046/j.1365-2958.2002.02984.x; pmid: 12100551 24. A. C. Teo, D. I. Roper, Core steps of membrane-bound peptidoglycan biosynthesis: Recent advances, insight and opportunities. Antibiotics 4, 495–520 (2015). doi: 10.3390/ antibiotics4040495; pmid: 27025638 25. B. C. Chung et al., Crystal structure of MraY, an essential membrane enzyme for bacterial cell wall synthesis. Science Orta et al., Science 381, eadg9091 (2023) 14 July 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E 341, 1012–1016 (2013). doi: 10.1126/science.1236501; pmid: 23990562 5760–5767 (2010). doi: 10.1016/j.vaccine.2010.06.087; pmid: 20619379 26. E. H. Mashalidis et al., Chemical logic of MraY inhibition by antibacterial nucleoside natural products. Nat. Commun. 10, 2917 (2019). doi: 10.1038/s41467-019-10957-9; pmid: 31266949 27. J. K. Hakulinen et al., MraY-antibiotic complex reveals details of tunicamycin mode of action. Nat. Chem. Biol. 13, 265–267 (2017). doi: 10.1038/nchembio.2270; pmid: 28068312 28. B. Al-Dabbagh et al., Active site mapping of MraY, a member of the polyprenyl-phosphate N-acetylhexosamine 1-phosphate transferase superfamily, catalyzing the first membrane step of peptidoglycan biosynthesis. Biochemistry 47, 8919–8928 (2008). doi: 10.1021/bi8006274; pmid: 18672909 29. V. V. Mokhonov et al., SlyD-deficient Escherichia coli strains: A highway to contaminant-free protein extraction. Biochem. Biophys. Res. Commun. 499, 967–972 (2018). doi: 10.1016/ j.bbrc.2018.04.029; pmid: 29626483 30. C. Löw et al., Crystal structure determination and functional characterization of the metallochaperone SlyD from Thermus thermophilus. J. Mol. Biol. 398, 375–390 (2010). doi: 10.1016/ j.jmb.2010.03.014; pmid: 20230833 31. U. Weininger et al., NMR solution structure of SlyD from Escherichia coli: Spatial separation of prolyl isomerase and chaperone function. J. Mol. Biol. 387, 295–305 (2009). doi: 10.1016/j.jmb.2009.01.034; pmid: 19356587 32. E. M. Quistgaard et al., Molecular insights into substrate recognition and catalytic mechanism of the chaperone and FKBP peptidyl-prolyl isomerase SlyD. BMC Biol. 14, 82 (2016). doi: 10.1186/s12915-016-0300-3; pmid: 27664121 33. S. Pazicky, A. A. Werle, J. Lei, C. Löw, U. Weininger, Impact of distant peptide substrate residues on enzymatic activity of SlyD. Cell. Mol. Life Sci. 79, 138 (2022). doi: 10.1007/s00018- 022-04179-4; pmid: 35184231 34. L. Martino et al., The interaction of the Escherichia coli protein SlyD with nickel ions illuminates the mechanism of regulation of its peptidyl-prolyl isomerase activity. FEBS J. 276, 4529–4544 (2009). doi: 10.1111/j.1742-4658.2009.07159.x; pmid: 19645725 35. Y. Zheng, D. K. Struck, T. G. Bernhardt, R. Young, Genetic analysis of MraY inhibition by the phiX174 protein E. Genetics 180, 1459–1466 (2008). doi: 10.1534/genetics.108.093443; pmid: 18791230 36. Y. Zheng, D. K. Struck, R. Young, Purification and functional characterization of phiX174 lysis protein E. Biochemistry 48, 4999–5006 (2009). doi: 10.1021/bi900469g; pmid: 19379010 37. B. C. Chung et al., Structural insights into inhibition of lipid I production in bacterial cell wall synthesis. Nature 533, 557–560 (2016). doi: 10.1038/nature17636; pmid: 27088606 38. J. Jumper et al., Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021). doi: 10.1038/ s41586-021-03819-2; pmid: 34265844 39. P. Kudela, V. J. Koller, W. Lubitz, Bacterial ghosts (BGs)— Advanced antigen and drug delivery system. Vaccine 28, 40. U. B. Mayr et al., Bacterial ghosts as antigen delivery vehicles. Adv. Drug Deliv. Rev. 57, 1381–1391 (2005). doi: 10.1016/ j.addr.2005.01.027; pmid: 15878634 41. G. Halfmann, F. Götz, W. Lubitz, Expression of bacteriophage PhiX174 lysis gene E in Staphylococcus carnosus TM300. FEMS Microbiol. Lett. 108, 139–143 (1993). doi: 10.1111/ j.1574-6968.1993.tb06089.x; pmid: 8486239 42. E. Henrich et al., Lipid Requirements for the Enzymatic Activity of MraY Translocases and in Vitro Reconstitution of the Lipid II Synthesis Pathway. J. Biol. Chem. 291, 2535–2546 (2016). doi: 10.1074/jbc.M115.664292; pmid: 26620564 43. E. Henrich et al., Analyzing native membrane protein assembly in nanodiscs by combined non-covalent mass spectrometry and synthetic biology. eLife 6, e20954 (2017). doi: 10.7554/ eLife.20954; pmid: 28067619 44. A. O. Oluwole et al., Peptidoglycan biosynthesis is driven by lipid transfer along enzyme-substrate affinity gradients. Nat. Commun. 13, 2278 (2022). doi: 10.1038/s41467-022- 29836-x; pmid: 35477938 45. W. D. Roof, R. Young, F X174 lysis requires slyD, a host gene which is related to the FKBP family of peptidyl- prolyl cis-trans isomerases. FEMS Microbiol. Rev. 17, 213–218 (1995). doi: 10.1016/0168-6445(94)00073-5; pmid: 7669348 46. T. Ikura, N. Ito, Requirements for peptidyl-prolyl isomerization activity: A comprehensive mutational analysis of the substrate-binding cavity of FK506-binding protein 12. Protein Sci. 16, 2618–2625 (2007). doi: 10.1110/ps.073203707; pmid: 18029417 47. J. C. Fiddes, G. N. Godson, Evolution of the three overlapping gene systems in G4 and phi X174. J. Mol. Biol. 133, 19–43 (1979). doi: 10.1016/0022-2836(79)90249-3; pmid: 160951 48. A. Pavesi, Origin, evolution and stability of overlapping genes in viruses: A systematic review. Genes 12, 809 (2021). doi: 10.3390/genes12060809; pmid: 34073395 49. K. R. Chamakura et al., Rapid de novo evolution of lysis genes in single-stranded RNA phages. Nat. Commun. 11, 6009 (2020). doi: 10.1038/s41467-020-19860-0; pmid: 33243984 50. A. Punjani, J. L. Rubinstein, D. J. Fleet, M. A. Brubaker, cryoSPARC: Algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296 (2017). doi: 10.1038/nmeth.4169; pmid: 28165473 51. A. Rohou, N. Grigorieff, CTFFIND4: Fast and accurate defocus estimation from electron micrographs. J. Struct. Biol. 192, 216–221 (2015). doi: 10.1016/j.jsb.2015.08.008; pmid: 26278980 53. E. F. Pettersen et al., UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2021). doi: 10.1002/pro.3943; pmid: 32881101 54. A. M. Waterhouse, J. B. Procter, D. M. Martin, M. Clamp, G. J. Barton, Jalview Version 2—A multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191 (2009). doi: 10.1093/bioinformatics/btp033; pmid: 19151095 55. M. Steinegger, J. Söding, Clustering huge protein sequence sets in linear time. Nat. Commun. 9, 2542 (2018). doi: 10.1038/ s41467-018-04964-5; pmid: 29959318 AC KNOWLED GME NTS We are particularly grateful to R. Young and T. Bernhardt for providing inspiration and feedback during the course of this project. We thank them, M. Kurosu, and D. Rees for critical feedback on the manuscript. We further thank R. Young for providing the DslyD strain. (Cryo) Electron microscopy was done in the Beckman Institute Resource Center for Transmission Electron Microscopy at Caltech. We are grateful to S. Chen for help with data collection and processing. Funding: This work was funded by the following: National Institutes of Health grant R01GM114611 (to W.M.C. and M.K.); National Institutes of Health grant DP1GM105385 (to W.M.C.); and the G. Harold and Leila Y. Mathers Foundation (to W.M.C.) Author contributions: Conceptualization: S.T. and W.M.C. Methodology: A.K.O., N.R., Y.E.L., S.T., H.G.Y., L.K., and W.M.C. Investigation: A.K.O., N.R., Y.E.L., S.T., H.G.Y., L.K., W.M.C. Visualization: A.K.O., N.R., S.T., and W.M.C. Funding acquisition: W.M.C. Project administration: W.M.C. Supervision: W.M.C. Writing – original draft: A.K.O., N.R., and W.M.C. Writing – review and editing: A.K.O., N.R., Y.E.L., S.T., H.G.Y., L.K., and W.M.C. Competing interests: Authors declare that they have no competing interests. Data and materials availability: All experimental data are available in the main text or the supplementary materials. Coordinates with experimental maps have been deposited to the RCSB or the EMDB for both the YESID21 (PDB ID 8G01, EMDB-29641) and YESFX174 (PDB ID 8G02 and EMDB- 29642) complexes. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg9091 Figs. S1 to S14 Table S1 References (56–57) MDAR Reproducibility Checklist Movie S1 52. R. Sanchez-Garcia et al., DeepEMhancer: A deep learning View/request a protocol for this paper from Bio-protocol. solution for cryo-EM volume post-processing. Commun. Biol. 4, 874 (2021). doi: 10.1038/s42003-021-02399-1; pmid: 34267316 Submitted 4 February 2023; accepted 23 May 2023 10.1126/science.adg9091 Orta et al., Science 381, eadg9091 (2023) 14 July 2023 8 of 8
10.1126_science.adh0993
RES EARCH PLASTIC UPCYCLING Chemical upcycling of polyethylene, polypropylene, and mixtures to high-value surfactants Zhen Xu1†, Nuwayo Eric Munyaneza1†, Qikun Zhang2, Mengqi Sun1, Carlos Posada1, Paul Venturo1, Nicholas A. Rorrer3,4, Joel Miscall3,4, Bobby G. Sumpter5*, Guoliang Liu1,6* Conversion of plastic wastes to fatty acids is an attractive means to supplement the sourcing of these high-value, high-volume chemicals. We report a method for transforming polyethylene (PE) and polypropylene (PP) at ~80% conversion to fatty acids with number-average molar masses of up to ~700 and 670 daltons, respectively. The process is applicable to municipal PE and PP wastes and their mixtures. Temperature-gradient thermolysis is the key to controllably degrading PE and PP into waxes and inhibiting the production of small molecules. The waxes are upcycled to fatty acids by oxidation over manganese stearate and subsequent processing. PP b-scission produces more olefin wax and yields higher acid-number fatty acids than does PE b-scission. We further convert the fatty acids to high-value, large–market-volume surfactants. Industrial-scale technoeconomic analysis suggests economic viability without the need for subsidies. A s the two most widely used commodity plastics, polyethylene (PE) (Table 1) and polypropylene (PP) contribute nearly 60% of the world’s plastic production (~400 million tonnes), primarily for short- term applications (1). The manufacturing of PE and PP is associated with the highest en- ergy consumption among all plastics (1, 2) and contributes substantially to annual greenhouse gas emissions (2). Short-term use plastics quickly turn into waste and cause substantial pollution (3). To recycle PE and PP, the waste collection and sorting processes must be economically efficient to lower the cost (4), and the recycled products should ideally be high value and high volume to have a major impact on waste ac- cumulation. Although PE and PP can be sepa- rated from heavier-than-water polymers such as polyvinyl chloride (PVC) and polyethylene ter- ephthalate (PET) through a sink-float method that uses water as the medium (Fig. 1A), further separation of PE and PP is much more chal- lenging because of their similar structures and densities. The two polymers are furthermore incompatible and cannot be blended unless expensive and sophisticated compatibilizers are used (5). Finding a generic and profitable method to recycle or upcycle both PE and PP while increasing their final product value over that of virgin plastics is thus imperative (3, 6–8). 1Department of Chemistry, Virginia Tech, Blacksburg, VA 24061, USA. 2Department of Chemistry, Chemical Engineering and Materials Science, Ministry of Education Key Laboratory of Molecular and Nano Probes, Shandong Normal University, Shandong 250014, PR China. 3Renewable Resources and Enabling Sciences Center, National Renewable Energy Laboratory, Golden, CO 80401, USA. 4BOTTLE Consortium, Golden, CO 80401, USA. 5Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA. 6Department of Chemical Engineering, Department of Materials Science and Engineering, Macromolecules Innovation Institute, Virginia Tech, Blacksburg, VA 24061, USA. *Corresponding author. Email: gliu1@vt.edu (G.L.); sumpterbg@ornl.gov (B.G.S.) †These authors contributed equally to this work. Chemical upcycling increases product value and is envisioned as a solution that converts postconsumer wastes into high-value chem- icals, e.g., the conversion of polystyrene (PS) into aryl carbonyls and aryl alkyls (9–11). Chem- ical upcycling of PP and PE, however, is diffi- cult because of the high ceiling temperatures involved (12). Moreover, the lack of heteroatom- associated weak links within the polymer chains (e.g., esters in PET) provides no selective chain- scission sites. Product control is thus exception- ally challenging. Recently, chemical reactions that use iridium- (13) and platinum-based cat- alysts (14) and ionic liquids (15) generated fuels and improved selectivity toward aryl moieties in the upcycling of PE. On a laboratory scale, PE can also be converted to propylene by dehydro- genation and tandem isomerizing ethenolysis (16, 17). Considering the technoeconomic po- tential for increased product value and gener- alizability to both PE and PP, the conversion of polyolefins to high-value fatty acids or ionic surfactants is appealing because PE and PP are aliphatic by nature. In addition, surfactant products have huge market demands (i.e., com- parable to plastics) and high economic values (i.e., higher than fuels, waxes, and regular aro- matic compounds) (table S1). Moreover, sur- factant products such as soaps are made from fatty acids of varying chain lengths and are often blended with ketones and aldehydes to soften the product and modulate fragrance release (18), diminishing the need to remove ketone and aldehyde by-products in the up- cycling process of PE and PP. Biodegradation converts PE to fatty acids by means of micro- bial strains, but the fermentation period is too long (>10 days) to be practically deployable in the chemical industry (19). Chemically, the Novoloop method and hydrothermal reactions with strong oxidants quickly degrade PE but require harsh reaction conditions, such as strong nitric acid and high pressure (20, 21). Most re- cently, metallization of PE over Zr and alkyl aluminum has afforded products with con- trollable average carbon-chain lengths, and the catalysts must operate under O2-, CO2-, and H2O-free conditions (22). Therefore, time- and material-efficient methods that use noncorro- sive chemicals, atmospheric pressure, and air- tolerant reaction conditions are highly sought to efficiently transform both PE and PP into large–market-volume fatty acids, preferably with- out the need for additional sorting and sepa- ration but with high selectivity and conversion. Here we report a gradient-temperature ther- molysis method that can selectively break PE, PP, and their mixtures into waxes under atmo- spheric pressure (Fig. 1, A and B). The key is that the temperature gradient prevents violent pyrol- ysis reactions, quenches vaporized waxes, and inhibits complete degradation to small mole- cules (Fig. 1C and figs. S1 and S2; see discussion in supplementary materials). PE- and PP-derived waxes are subsequently transformed into fat- ty acids with high acid numbers (ANs) and number-average molar masses of up to ~700 and 670 Da, respectively (Fig. 1D and tables S2 to S4). Through subsequent saponifica- tion, we obtain an ionic surfactant product that contains salts of fatty acids. Simply mix- ing with additives (e.g., fragrances) can produce commercial products, such as soap bars and liquid detergents (examples in Fig. 1A), which have higher market values than typical chemical products such as fuels and alkylaromatics (11, 14). Polyethylene conversion to fatty acids Initially, pulverized lab-grade PE (~500 mg; Mw, 97 kDa; polydispersity Ð = 2) was loaded into a custom-designed quartz reactor (Fig. 1A and fig. S1) and purged with gases of con- trolled compositions (N2, 10 vol % O2 in N2, or air). The degradation was initiated by heating the bottom of the reactor to ~360°C stepwise with a step size of ~100°C per 5 min. Polymer smokes (fig. S2) appeared, indicating the va- porization of fragmented polyolefins or waxes. The waxes were condensed in the cold part of the reactor, preventing further fragmentation to shorter hydrocarbons. The wax yields from PE degradation in N2 (PE-N2-wax), 10 vol % O2 (PE-O-wax), or air (PE-air-wax) were 86, 79, and 55 wt % (Fig. 2A and table S2, experiments 1 to 3, 7 to 9, and 13), respectively. In a control experiment under N2 without a temperature gradient, the products were mostly short-chain hydrocarbons of C8 and below (Fig. 1C). Gas chromatography–mass spectrometry (GC-MS) was used to characterize the wax composition (Fig. 2B), showing mainly solid waxes with minor fractions of light hydrocar- bons (~10 wt % of C8 and below, fig. S3). Each primary peak could be resolved into a doublet of alkene and alkane with the same carbon number (figs. S4 and S5), similarly to degrada- tion in flow reactors (23). In the presence of O2, Xu et al., Science 381, 666–671 (2023) 11 August 2023 1 of 6 ~630 and 540 (fig. S6 and table S3), corre- sponding to a carbon number of ~45 and 42, respectively. Despite m/z ranges of 300 to 1000 for PE-N2-wax and 200 to 900 for PE-O-wax (fig. S6C), the polydispersities were relatively small (<1.1, table S3). The PE-N2-wax and PE-O-wax were further characterized by nuclear magnetic resonance (NMR) spectroscopy through het- eronuclear multiple bond correlation (HMBC) and heteronuclear single quantum correlation (HSQC) experiments (Fig. 2C and figs. S7 and S8) to investigate the waxes’ structures. NMR confirmed the presence of unsaturated carbon in both PE-N2-wax and PE-O-wax, showing pri- marily 2-propenyl at the chain end (1H NMR d 5.0 and 5.8, 13C NMR d 114 and 138) and minor internal alkenes (13C NMR d 123 to 131). The presence of O2 partially oxidized PE-O-wax and produced ketones, aldehydes, and esters (fig. S8). Upcycling of the waxes was conducted over Mn compounds for 10 hours in an airflow at 150°C. Although inorganic MnO2 and KMnO4 have shown effectiveness in paraffin oxidation (28), they were ineffective in oxidizing PE-derived wax, probably because of low miscibility, showing no appreciable carbonyl signals after 24 hours (fig. S9). By contrast, Mn stearate catalyzed the oxidization of PE-derived wax much faster owing to the better dispersion of the catalyst in the organic media (29, 30). The oxidation rates of PE-N2-wax or PE-O-wax were similar over Mn stearate (5 wt % loading) un- der a constant air flow, as shown by the similar slopes of carbonyl index (CI) changes as a function of time. PE-O-wax showed a higher final CI because of the higher initial value than did PE-N2-wax. We conclude on the basis of HMBC analysis that the oxidation of PE-O-wax over Mn stearate intensified the carbonyl con- centration in the first 6 hours (figs. S9 and S10, 13C NMR d 160 to 206), producing primarily aldehydes and minor esters, ketones, and car- boxylic acids. The oxidation likely occurred by means of a radical addition mechanism (31, 32), possibly through epoxy intermediate rear- rangement to aldehydes and ketones (33). The presence of aldehyde was confirmed by NMR- HMBC, whereas epoxy signals were too weak to D RES EARCH | R E S E A R C H A R T I C L E PE-O-wax and PE-air-wax exhibited GC-MS peaks and molecular ion signals similar to those of PE-N2-wax. PE-O-wax and PE-air-wax showed stronger intensities at shorter elution times than PE-N2-wax (Fig. 2B), indicating more short hydrocarbons and lower average molecu- lar weights as a result of accelerated degrada- tion by oxygen radicals (24). The PE-air-wax yield was too low (~55%), albeit higher than those in the literature (25, 26), to be practically useful for generating hydrocarbons suitable for downstream production of surfactants (Fig. 2A); therefore, no further characterization of PE-air-wax was conducted. High-temperature gel permeation chromatography (HT-GPC) con- firmed the degradation of polymers (fig. S6). Because HT-GPC cannot resolve the exact mo- lecular weights in this range, and because GC cannot detect low-volatility heavy hydrocar- bons (>C40) (27), atmospheric pressure chem- ical ionization mass spectrometry (APCI-MS) was used to analyze the full composition. The spectra of PE-N2-wax and PE-O-wax showed waxes of mass/charge ratio (m/z) centered at A B C Fig. 1. Upcycling of PE and PP to fatty acids in a temperature-gradient reactor. (A) Schematic process flow of separation and upcycling of commercial polyolefins, including high-density PE (HDPE), low- density PE (LDPE), and PP to soap products performed using a custom-designed gradi- ent thermal reactor. The temperature- gradient reactor has a hot and cold zone, preventing complete thermolysis of PE and PP to small molecules, and is key to controlling the chain length of fragmented products. Photographs show representative PE and PP wastes used in this study (HDPE container, grocery bags, sandwich bags, bottle caps, PP centrifuge tubes, and PP foam), as well as the products of intermediate waxes, fatty acids, surfactant solution, and soap molded into various shapes. The market price per metric ton of common surfactant products is almost double that of virgin plastics (table S1). (B) Reaction schemes of upcycling PE and PP to fatty acids. (C) Representative product distributions after PE degradation in the reactor with and without temperature gradient. The distribution and intensity were measured with GC. Signals from 3 to 7 min overlapped with the toluene solvent and thus were not shown for clarity. (Inset) Infrared thermal image of the reactor with and without temperature gradient. (D) AN of resulting fatty acids compared with that of stearic acid (SA, C18) and theoretical values of fatty acids with average carbon numbers of C47 for PE and C45 for PP. The error bars are the standard deviation of at least three replicates. Xu et al., Science 381, 666–671 (2023) 11 August 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E be confirmed. The dominance of aldehyde in- termediates agreed well with the simulation results. Because aldehydes were susceptible to further oxidation, they were converted to acids (Fig. 2D, 13C NMR d 180) in the latter 2 hours of oxidation and subsequent saponification in 0.1 M aqueous KOH. The saponification hydrolyzed minor fatty esters and increased the amount of fatty acids; it also helped remove undesired short-chain acids (e.g., acetic acids). Hydroxyl signals were too weak to confirm the presence of alcohols in the product. The saponified solution was neutralized with HCl, crashing out fatty acid (PE-O-FA). After washing and drying under a vacuum, PE-O-FA was char- acterized by HMBC, showing primarily carbox- ylic acids and minor ketones. Aldehydes were no longer detectable (fig. S11). The carboxyl in the fatty acid correlated with two types of pro- tons (Fig. 2D, 1H NMR d 2.0 and 1.5), poten- tially indicating a carboxyethyl end group. The AN of the PE-O-FA was determined by tit- ration, giving an ordinary AN of 96 mg KOH/g (Fig. 1D and table S4) that slightly exceeded the theoretical value of ~80 mg KOH/g for a 700 g/ mol monocarboxylic acid (C47). Extension to polypropylene and mixtures After successfully converting PE to fatty acids, the process was applied to pulverized lab-grade PP (500 mg; Mw, 158 kDa; Ð=4) under reaction conditions similar to those used for PE (N2, 10 vol % O2 in N2, and air). Degradation products in N2 (PP-N2-wax), 10 vol % O2 (PP-O-wax), and air (PP-air-wax) showed total wax yields of 90, 85, and 87 wt %, respectively (Fig. 3A and table S2, experiments 4 to 6, 10 to 12, and 14), with a small amount of coke and gaseous products (fig. S12). The m/z ranges were 170 to 950 (cen- tered at 560) for PP-N2-wax and 150 to 700 (cen- tered at 450) for PP-air-wax (fig. S6C), and the molecular weight polydispersities were rela- tively small (table S3). Unlike that for PE, the PP wax yield remained high in the air, eliminat- ing the need for controlled gases in the process and making the method more economically attractive. Therefore, the degradation of PP in 10 vol % O2 was not investigated further. GC-MS analysis of PP-N2-wax showed broad multimodal distributions of primarily alkene products in GC-MS (fig. S13). The carbon num- bers of alkenes were mostly multiples of 3 (or 3n in the range of C9 to C36), and those of the rest of alkanes, alkenes, and dienes were main- ly 3n+1. This distribution suggested that the primary degradation mechanism could be chain scission on the PP backbone because each PP repeating unit has three carbons (which is in agreement with our simulations below). With air present, the chromatogram of PP-air-wax became crowded and hard to analyze (fig. S6A). PP-N2-wax and PP-air-wax were characterized by APCI-MS (fig. S6, B and C), and the latter showed a slightly higher proportion of short Table 1. Abbreviations. Full name Abbreviation Chemicals ..................................................................................................................................................................................................................... Polyethylene ..................................................................................................................................................................................................................... Polypropylene ..................................................................................................................................................................................................................... Polyvinyl chloride ..................................................................................................................................................................................................................... Polyethylene terephthalate ..................................................................................................................................................................................................................... High-density PE ..................................................................................................................................................................................................................... Low-density PE ..................................................................................................................................................................................................................... Stearic acid ..................................................................................................................................................................................................................... Wax yielded from PE degradation in N2 ..................................................................................................................................................................................................................... Wax yielded from PE degradation in 10 vol % O2 ..................................................................................................................................................................................................................... Wax yielded from PE degradation in air ..................................................................................................................................................................................................................... PE-O-wax derived fatty acid ..................................................................................................................................................................................................................... Wax yielded from PP degradation in N2 ..................................................................................................................................................................................................................... Wax yielded from PP degradation in 10 vol % O2 ..................................................................................................................................................................................................................... Wax yielded from PP degradation in air ..................................................................................................................................................................................................................... PP-air-wax derived fatty acid ..................................................................................................................................................................................................................... PE PP PVC, or V PET HDPE LDPE SA PE-N2-wax PE-O-wax PE-air-wax PE-O-FA PP-N2-wax PP-O-wax PP-air-wax PP-air-FA Instrumentation and methods ..................................................................................................................................................................................................................... Gas chromatography ..................................................................................................................................................................................................................... Gas chromatography-mass spectrometry ..................................................................................................................................................................................................................... High-temperature gel permeation chromatography ..................................................................................................................................................................................................................... Atmospheric pressure chemical ionization mass spectrometry ..................................................................................................................................................................................................................... Nuclear magnetic resonance spectroscopy ..................................................................................................................................................................................................................... Heteronuclear multiple bond correlation ..................................................................................................................................................................................................................... Heteronuclear single quantum correlation ..................................................................................................................................................................................................................... Proton NMR ..................................................................................................................................................................................................................... Quantitative NMR ..................................................................................................................................................................................................................... Flame ionization detector ..................................................................................................................................................................................................................... Mass selective detector ..................................................................................................................................................................................................................... Density functional–based tight-binding ..................................................................................................................................................................................................................... ab initio molecular dynamics ..................................................................................................................................................................................................................... Fourier transform infrared spectroscopy ..................................................................................................................................................................................................................... Thermogravimetric analysis ..................................................................................................................................................................................................................... GC GC-MS HT-GPC APCI-MS NMR HMBC HSQC 1H NMR q-NMR FID MSD DFTB AIMD FTIR TGA Parameters ..................................................................................................................................................................................................................... Number-average molar mass ..................................................................................................................................................................................................................... Weight-average molar mass ..................................................................................................................................................................................................................... Polydispersity ..................................................................................................................................................................................................................... Carbonyl index ..................................................................................................................................................................................................................... Acid number ..................................................................................................................................................................................................................... Activation energy ..................................................................................................................................................................................................................... Pre-exponential factor ..................................................................................................................................................................................................................... Selectivity of thermolysis (T) or upcycling (U) ..................................................................................................................................................................................................................... Yield of thermolysis (T) or upcycling (U) ..................................................................................................................................................................................................................... Concentration of alkenyl groups ..................................................................................................................................................................................................................... Concentration of total waxes ..................................................................................................................................................................................................................... Concentration (mmol/g) of acid group in fatty acids ..................................................................................................................................................................................................................... Concentrations (mmol/g) of alkenyl groups in fatty acids ..................................................................................................................................................................................................................... Mn Mw Ð CI AN Ea A sT or sU aT or aU cc¼c cwax cacid cc¼c;FA hydrocarbons because of air-induced oxidation, which is similar to the finding in a previous re- port. (34) As with the PE degradation product, PP-N2-wax and PP-air-wax were primarily com- posed of terminal alkenes with minor internal alkenes (Fig. 3B and figs. S14 and S15). The pri- mary form of the terminal alkene was possibly 2-methyl-2-propenyl, as suggested by the cor- relations in the HMBC and HSQC. Additionally, some minor forms of alkenes were detected in the 13C NMR spectra and could be ascribed to internal alkenyl structures (fig. S14A). PP-air- wax contained ketones, aldehydes, and esters because of partial oxidation (fig. S15), but the alkenyl region resembled that of the PP-N2-wax. As with the PE-derived waxes, upcycling of PP-N2-wax and PP-air-wax was conducted in air Xu et al., Science 381, 666–671 (2023) 11 August 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E A B C Fig. 2. Degradation of PE into intermediate waxes and upcycling into fatty acids. (A) Yields and (B) GC-MS chromatograms of intermediate waxes after degrading lab-grade PE in N2, 10 vol % of O2 balanced with N2, or air. (C) NMR-HMBC spectra of PE-N2-wax in deuterated p-xylene. The circles highlight the C-H correlations of the major and minor alkene products. (D) NMR-HMBC spectrum of the fatty acid derived from PE-O-wax. The black box highlights the two carboxyl carbon correlations with protons. [Insets in (C) and (D)] Chemical structures of the major alkene and fatty acid products. at 150°C over Mn stearate. PP-air-wax oxition was faster than that for PP-N2-wax; the former showed a CI of ~0.9 after 4 hours (fig. S9C). Moreover, a possible epoxy structure was de- tected in oxidized PP-air-wax (fig. S16), indi- cating that poxy could be a potential oxidation intermediate between alkenyl and aldehydes. The fatty acids (PP-air-FA) that resulted from hydrolysis were characterized by NMR-HMBC and showed no detectable aldehyde but stronger ketone carbonyl signals than those of PE-O-FA. The acid carbonyl showed three correlations with protons (Fig. 3C and fig. S17), possibly indicating a 2-carboxylpropyl terminal structure. PP-air-FA exhibited an AN of 169 mg KOH/g, substantially higher than the theoretical value of 84 mg KOH/g for a monocarboxylic acid at 670 g/mol (C46) (table S4, entry 6), suggesting the presence of polyacids. After both PE and PP were successfully con- verted to wax and fatty acids at high conversions, municipal PE/PP wastes (obtained in Blacksburg, VA) were sorted (if labeled), pulverized, and mixed to mimic a waste stream [high-density PE (HDPE, 25 wt %), PP (25 wt %), low-density PE (LDPE, 25 wt %), and PE/PP unsorted (25 wt %)]. Three batches of the PE/PP waste mixtures (500 mg per batch) were degraded into wax, resulting in an average yield of 82% in N2. The lower wax yield of PE/PP mixtures from municipal waste streams was ascribed to the additives, paint, dye, and labels. The wax was further upcycled to fatty acids, giving an aver- age AN of 57.3 mg KOH/g (Fig. 1D and table S4, entry 8). The AN can be improved by tun- ing the oxygen level during degradation, the airflow during oxidation, and the upcycling temperature. Ketones and aldehydes could potentially be present in the product and can adjust the product viscosity, function as soap softeners (35), and modulate fragrance release (18). Theoretical simulations Density functional–based tight-binding (DFTB) simulations were used to gain a molecular understanding of the temperature-gradient degradation of PE and PP and to evaluate the subsequent oxidative upcycling reactions. DFTB and the extended tight-binding methods enable simulations of relatively large systems and rea- sonable timescales with good accuracy but are considerably faster than typical ab initio den- sity functional theory (DFT) methods (36). Fig. 3. Degradation of PP and PE/PP mixture into intermediate waxes and upcycling into fatty acids. (A) The intermediate wax yield of lab-grade PP upon degradation in N2, N2/O2 mixture (10 vol % O2), or air, along with the wax yield from a PE/PP waste mixture (including 25 wt % HDPE, 25 wt % PP, 25 wt % LDPE, and 25 wt % PE/PP unsorted) upon degradation in N2. The ANs are shown in table S4. The low wax yield from the PE/PP waste mixture was due to solid contaminants such as paper labels, paints, and dyes. (B) NMR- HMBC spectra of PP-N2-wax in deuterated p-xylene. (Inset) Chemical structure of the major alkene product. The circles highlight the C-H correlations of the major and minor alkene products, respec- tively. (C) NMR-HMBC spectra of PP-air-FA. (Inset) Chemical structure of the major fatty acid product. The black box highlights the three carboxyl carbon correlations with protons. Multiple (>10) molecular-dynamics trajectories were used to evaluate the PP and PE chain be- havior. The model polymer chains undergo sig- nificant coiling and dynamical oscillations during the ab initio molecular dynamics (AIMD) simulations. The mechanisms of alkene for- mation from PE and PP are primarily radical b-scission and secondarily disproportionation (Fig. 4A and fig. S18), similar to the findings of previous reports (3, 7, 38). We focus the discus- sion on PP because it produces more olefin wax Xu et al., Science 381, 666–671 (2023) 11 August 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Plausible reaction pathway, thermody- namics, and kinetics of b-chain scission in PE and PP. (A) The thermolysis of a model polymer chain into an alkane and alkene fragment and subsequent oxidation to a fatty acid, as captured by the simulations. (B) b-Scission of a radical in PE or PP, R=H or CH3. (C) Thermodynamic parame- ters and (D) kinetic constants of PE b-scission and PP b-scission at 298, 633, and 823 K, as calculated according to the Benson group-additivity method (39, 40). during degradation than does PE. The results show that chain scission occurs somewhat ran- domly along the backbone of the PP chain and that multiple bonds are broken to form alkene and alkane fragments (Fig. 4A and fig. S19). We note that the simulated fragment-size distribu- tions agree with experimental observation and that there are more alkenes formed for PP than for PE. For example, in the experimental PE degradation, the initial C=C concentration in PE-N2-wax was ~1.15 mmol/g on the basis of quantitative NMR (q-NMR, fig. S20 and table S5). By contrast, the C=C concentration in PP- N2-wax was ~3.15 mmol/g, showing good qualitative agreement with the simulation results. The degradation fragments from the simu- lations were then annealed by geometry op- timization and modeled for reactivity toward oxygen at reduced temperatures. During the oxidation step, hydrogen extraction from a methyl or methylene group in the fragment was predicted to form a hydroperoxyl radical, which is prone to react with a readily formed aldehyde from alkenyl oxidation at the frag- ment end (Fig. 4A). This sequence initially produces an alcohol-like intermediate (H is transferred to carbonyl O on the aldehyde) that ultimately forms a carboxylic acid, by using a modified Baeyer-Villiger reaction. After the al- dehyde accepts a H from the HOO(cid:129), the O-O(cid:129) binds to the intermediate C-OH to form a car- boxylic acid. The aldehyde intermediate was confirmed by the NMR-HMBC experiments on partially oxidized wax (fig. S16). The primary intermediates found in this work also agree with the results of a previous study of paraffin oxidation over stearate (29), but we cannot rule out other potential oxidation pathways at this time. Applying the qualitative insights from the DFTB simulations, we then used the Benson group-additivity method (39, 40) to evaluate the standard Gibbs free energy (DGo) (Fig. 4B and table S6) of PE b-scission (DGo PE = 51.1 kJ/mol) and PP b-scission (DGo PP = 45.7 kJ/mol). Thermo- dynamically, alkene formation from PP and PE is disfavored at 298 K (Fig. 4C and table S7). At elevated temperatures (633 K and 823 K), PE b-scission and PP b-scission become increas- ingly thermodynamically favorable. Under our experimental conditions (633 K), the reaction enthalpies and DGoof PE and PP degradation are reduced (DHPP;633K = 88.9 kJ/mol and DGPP;633K = −5.60 kJ/mol; DHPE;633K = 94.3 kJ/mol and DGPE;633K = 1.20 kJ/mol). However, the process is still slightly disfavored for PE. Upon increas- ing the temperature to 823 K, both PE and PP have favored b-scission pathways (DHPP;823K = 86.9 kJ/mol and DGPP;823K = −33.1 kJ/mol; DHPE;823K = 92.9 kJ/mol and DGPE;823K = −26.5 kJ/mol). The kinetic parameters and con- stants of b-scission were also predicted by using the group-additivity method on the basis of model reactions (fig. S21A) (41). PP b-scission showed lower activation energies and higher Xu et al., Science 381, 666–671 (2023) 11 August 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E kinetic constants than did PE at all tempera- tures (Fig. 4D and table S8). In general, alkene formation from PP b-scission is more favored and faster than that from PE (this agrees well with our experiments and simulations above). Our observation of PE– and PP–b-scission de- pendence on temperature also agrees well with the literature, in which the alkene yield from PE thermolysis increased from ~30% at ~600 K to ~70% at above 800 K, whereas that from PP thermolysis stayed high (>70%) regardless of the temperature (fig. S22) (23, 37, 42–48). Technoeconomic analysis (TEA) of the process was assessed on the basis of a 10,000-ton/year capacity (tables S9 to S14). The total capital investment was estimated at USD 2.70 million with an annual net profit of USD 1.03 million, resulting in an internal rate of return of 39.1%, a payback period of 2.63 years, and an aver- age return of investment of 33.0%, with- out any government subsidies or tax returns (table S15). Outlook The fatty acids produced in this study have a wide range of applications, either for being kept in the plastic loop or for downstream uses. The downstream surfactant products have at least twice the market value of virgin plastics, rep- resenting an economically competitive process for plastic-waste utilization. In addition, sur- factants have a market volume matching that of end-of-life plastic wastes, thus representing a volume-impactful method for plastic-waste re- moval. Controlled thermolysis in a temperature- gradient reactor is the key to controlling the high yield of the wax products relative to small gaseous molecules. Unlike existing processes (22), our process tolerates oxygen and requires no expensive catalysts or stringent reaction con- ditions. The resulting products show good ANs for PE- (96 mg KOH/g) and PP-derived fatty acids (169 mg KOH/g), which can be further optimized by inhibiting side reactions (figs. S23 and S24). We anticipate the process to be amenable to a diverse range of other plastic wastes for producing high-value, large–market- volume products (e.g., fatty alcohols and sul- fates) (49–51). RE FE RENCES AND N OT ES 3. J. Hopewell, R. Dvorak, E. Kosior, Philos. Trans. R. Soc. Lond. Ser. B 364, 2115–2126 (2009). 4. Y. Zhang, G. Wang, Q. Zhang, Y. Ji, H. Xu, Environ. Impact Assess. Rev. 93, 106728 (2022). 45. M. T. S. P. De Amorim, C. Comel, P. Vermande, J. Anal. Appl. Pyrolysis 4, 73–81 (1982). 46. J. Michal, J. Mitera, S. Tardon, Fire Mater. 1, 160–168 (1976). 47. G. Elordi, M. Olazar, G. Lopez, M. Artetxe, J. Bilbao, Ind. Eng. 5. B. A. Abel, R. L. Snyder, G. W. Coates, Science 373, 783–789 Chem. Res. 50, 6650–6659 (2011). (2021). 6. L. T. J. Korley, T. H. Epps III, B. A. Helms, A. J. Ryan, Science 373, 66–69 (2021). 7. C. Jehanno et al., Nature 603, 803–814 (2022). 8. H. Chen, K. Wan, Y. Zhang, Y. Wang, ChemSusChem 14, 4123–4136 (2021). 9. Z. Huang et al., J. Am. Chem. Soc. 144, 6532–6542 (2022). 10. R. Cao et al., Nat. Commun. 13, 4809–4819 (2022). 11. Z. Xu et al., Proc. Natl. Acad. Sci. U.S.A. 119, e2203346119 (2022). 12. R. Conley, Ed., Thermal Stability of Polymers (Marcel Dekker, 1970). 13. X. Jia, C. Qin, T. Friedberger, Z. Guan, Z. Huang, Sci. Adv. 2, e1501591 (2016). 14. F. Zhang et al., Science 370, 437–441 (2020). 15. W. Zhang et al., Science 379, 807–811 (2023). 16. R. J. Conk et al., Science 377, 1561–1566 (2022). 17. N. M. Wang et al., J. Am. Chem. Soc. 144, 18526–18531 2022). 18. A. Trachsel, C. Chapuis, A. Herrmann, Flavour Fragrance J. 28, 280–293 (2013). 19. S. Kumar Sen, S. Raut, J. Environ. Chem. Eng. 3, 462–473 (2015). 20. L. R. Melby, Macromolecules 11, 50–56 (2002). 21. J. Y. Yao et al., Methods for the Decomposition of Contaminated Plastic Waste. US Patent US 11,220,586 B2 (2012). 22. U. Kanbur et al., Chem 7, 1347–1362 (2021). 23. U. R. Gracida-Alvarez, M. K. Mitchell, J. C. Sacramento-Rivero, D. R. Shonnard, Ind. Eng. Chem. Res. 57, 1912–1923 (2018). 24. A. Holmström, E. Sörvik, J. Appl. Polym. Sci. 18, 3153–3178 (1974). 25. S. M. Al-Salem, A. Dutta, Ind. Eng. Chem. Res. 60, 8301–8309 (2021). 26. G. Celik et al., ACS Cent. Sci. 5, 1795–1803 (2019). 27. V. Vrkoslav, R. Míková, J. Cvačka, J. Mass Spectrom. 44, 101–110 (2009). 28. J. W. Frankenfeld, Ed., Study of Methods for Chemical Synthesis of Edible Fatty Acids and Lipids (NASA, 1968). 29. P. George, E. K. Rideal, A. Robertson, Proc. R. Soc. Lond. Ser. A 185, 288–309 (1946). 30. P. K. Roy, P. Surekha, R. Raman, C. Rajagopal, Polym. Degrad. Stabil. 94, 1033–1039 (2009). 31. F. Xiao, X. Sun, Z. Li, X. Li, ACS Omega 5, 12777–12788 (2020). 32. M. S. Stark, J. Am. Chem. Soc. 122, 4162–4170 (2000). 33. A. K. Pandey, R. K. Varshnaya, P. Banerjee, Eur. J. Org. Chem. 2017, 1647–1656 (2017). 34. K. P. Sullivan et al., Science 378, 207–211 (2022). 35. A. Meziani, D. Touraud, A. Zradba, M. Clausse, W. Kunz, J. Mol. Liq. 84, 301–311 (2000). 36. B. Hourahine et al., J. Chem. Phys. 152, 124101 (2020). 37. H. Bockhorn, A. Hornung, U. Hornung, D. Schawaller, J. Anal. Appl. Pyrolysis 48, 93–109 (1999). 38. T. Kuroki, T. Sawaguchi, S. Niikuni, T. Ikemura, Macromolecules 15, 1460–1464 (1982). 39. S. W. Benson, J. H. Buss, J. Chem. Phys. 29, 546–572 (1958). 40. R. L. Brown, S. E. Stein, in NIST Chemistry WebBook, NIST Standard Reference Database Number 69, P. J. Linstrom, W. G. Mallard, Eds. (National Institute of Standards and Technology, 2023); https://webbook.nist.gov/chemistry/. 41. M. K. Sabbe, M. F. Reyniers, V. Van Speybroeck, M. Waroquier, G. B. Marin, ChemPhysChem 9, 124–140 (2008). 42. A. Marcilla, M. I. Beltrán, R. Navarro, J. Anal. Appl. Pyrolysis 86, 14–21 (2009). 48. P. Das, P. Tiwari, Resour. Conserv. Recycling 128, 69–77 (2018). 49. H. Sabine, Oxidation of paraffinic hydrocarbons. US Patent US2391236A (1945). 50. D. V. Favis, Selective ozone oxidation of hydrocarbons. US Patent US2955123 (1956). 51. W. H. de Groot, Sulphonation Technology in the Detergent Industry (Springer, 1991). 52. Z. Xu et al., Chemical upcycling of polyethylene, polypropylene, and mixtures to high-value surfactants, Dryad (2023); https://doi.org/10.5061/dryad.c866t1gc5. AC KNOWLED GME NTS The authors acknowledge the Chemistry Chromatography Center at Virginia Tech, especially M. Ashraf-Khorassani, for providing GC and GC-MS support. The authors also acknowledge the Mass Spectrometry Incubator at Virginia Tech, especially R. Helm and K. Ray for their support on APCI-MS. The molecular simulations were performed at the Center for Nanophase Materials Sciences, a US Department of Energy (DOE) Office of Science User Facility operated at Oak Ridge National Laboratory. Funding: G.L. acknowledges support from the Department of Chemistry and the Dean’s Discovery Fund at Virginia Tech. This material is partially based on the work supported by NSF Award DMR-1752611 through the CAREER award. Mechanistic studies of the deconstruction processes were partially supported by the US DOE Office of Science, Materials Sciences and Engineering Division. The simulations used resources of the Compute and Data Environment for Science (CADES) at the Oak Ridge National Laboratory, which is supported by the Office of Science of the US DOE under contract no. DE-AC05-00OR22725. Funding for N.A.R. and J.M. was provided in part by the US DOE Office of Energy Efficiency and Renewable Energy, Advanced Manufacturing Office (AMO) and Bioenergy Technologies Office (BETO). This work was performed as part of the BOTTLE Consortium and was supported by AMO and BETO under contract no. DE-AC36-08GO28308 with the National Renewable Energy Laboratory, operated by Alliance for Sustainable Energy. Author contributions: G.L. conceived and supervised the project. G.L. and Z.X. designed the project. Z.X. and N.E.M. conducted degradation-upcycling reactions and GC, GC-MS, and FTIR measurements. Z.X. and P.V. performed NMR experiments. B.G.S. conducted DFT simulations. Q.Z. performed TEA and heat and mass transfer analyses. N.E.M. and C.P. performed rheological experiments. N.R. and J.M. conducted HT-GPC measurements. Z.X. determined the standard Gibbs free energies for the degradation reactions. M.S. performed titrations to determine the ANs. N.E.M. evaluated the effect of temperature gradient on the degradation reactions. G.L., B.G.S., Z.X., and N.E.M. analyzed the data, interpreted the results, and wrote the manuscript. All authors proofread the manuscript. Competing interests: G.L., Z.X., and N.E.M. are inventors on Provisional Patent Application no. 63/462,863 pertaining to this work. Data and materials availability: Data pertaining to this work are available in the supplementary materials and Dryad (52). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh0993 Figs. S1 to S31 Tables S1 to S18 References (53–103) 1. R. Geyer, J. R. Jambeck, K. L. Law, Sci. Adv. 3, e1700782 43. D. Zhao, X. Wang, J. B. Miller, G. W. Huber, ChemSusChem 13, (2017). 1764–1774 (2020). 2. S. R. Nicholson, N. A. Rorrer, A. C. Carpenter, G. T. Beckham, 44. J. K. Y. Kiang, P. C. Uden, J. C. W. Chien, Polym. Degrad. Stabil. Joule 5, 673–686 (2021). 2, 113–127 (1980). Submitted 9 February 2023; accepted 16 June 2023 10.1126/science.adh0993 Xu et al., Science 381, 666–671 (2023) 11 August 2023 6 of 6
10.1126_science.adg9500
RES EARCH QUANTUM SENSING Improving metrology with quantum scrambling Zeyang Li ( Roman Schmied5, Soonwon Choi4, Mikhail Lukin2, Edwin Pedrozo-Peñafiel1, Vladan Vuletić1* )1, Simone Colombo1, Chi Shu1,2, Gustavo Velez1,3, Saúl Pilatowsky-Cameo4, Quantum scrambling describes the spreading of information into many degrees of freedom in quantum systems, such that the information is no longer accessible locally but becomes distributed throughout the system. This idea can explain how quantum systems become classical and acquire a finite temperature, or how in black holes the information about the matter falling in is seemingly erased. We probe the exponential scrambling of a multiparticle system near a bistable point in phase space and utilize it for entanglement-enhanced metrology. A time-reversal protocol is used to observe a simultaneous exponential growth of both the metrological gain and the out-of-time-order correlator, thereby experimentally verifying the relation between quantum metrology and quantum information scrambling. Our results show that rapid scrambling dynamics capable of exponentially fast entanglement generation are useful for practical metrology, resulting in a 6.8(4)-decibel gain beyond the standard quantum limit. tum metrology, this enables a family of powerful quantum amplification protocols (17–24) such as signal amplification through time-reversed interaction (SATIN) (23). Such protocols can be robust against many limitations that usual- ly affect entanglement-enhanced atomic sen- sors, including imperfect measurements. In the presence of exponential scrambling dynamics (Fig. 1A), the SATIN signal is also amplified exponentially over time. Experimental setup The Lipkin-Meshkov-Glick (LMG) Hamiltonian (24, 25–34) is of particular interest in this con- text, as it exhibits exponential evolution in phase space while it can be experimentally realized in a cavity quantum electrodynamics (cQED) system. The latter is accomplished by adding a global rotation term ^S x to the one- axis-twisting (OAT) (35) Hamiltonian ^Sz 2 ð1Þ ^H ¼ c^Sz 2 þ W^S x (cid:3) (cid:1) Here S ¼ ^S x; ^S y; ^S z 2 represents the total spin of the system consisting of N ¼ 2S spin-1 particles, c is the shearing parameter for OAT, and W is the transverse rotation frequency. Although the time evolution is not chaotic because of the conservation of ^S 2 , the LMG Hamiltonian nevertheless features a quantum Lyapunov exponent for 0 < W= Scð Þ < 2 due to an unstable (bifurcating) trajectory in the system phase space (Fig. 1A) (32, 36, 37). Our experiments operate with N ¼ 200 171Yb atoms whose magnetic sublevels ↑; ↓j i in the 2 sys- electronic ground state represent a spin- 1 tem. One of the two spin states ( ↑j i) couples to an electronically excited state ej i via sþ-polarized light that circulates inside the optical cavity (Fig. 1B). The coupling between a single atom and the cavity is characterized by the single- atom cooperativity h ¼ 8:8 2ð Þ (38). We imple- ment the LMG Hamiltonian in the rotating frame by adding an oscillating transverse mag- netic field to the OAT Hamiltonian (34) [Fig. 1B and supplementary materials (SM) (39)]. The experiments start by initializing the system in a coherent spin state (CSS) pointing along the x axis by means of optical pumping followed by a p=2 spin rotation. Analytical solutions using the Holstein-Primakoff approx- Þ < 0 or imation (40) show that for W= Scð Þ > 2, the system evolution is periodic W= Scð (41). How- with a frequency w ¼ ever, for 0 < W= Scð Þ < 2 the frequency w becomes imaginary, corresponding to an ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi W2 (cid:1) 2ScW p E ven though all unitary dynamics of quan- tum systems are in principle reversible, it is extremely challenging in practice to reverse the arrow of time in generic in- teracting many-body systems. This is be- cause any small perturbations or imperfections in the time-reversed dynamics can lead to highly complicated, nonlocal changes in quan- tum wave functions, similar to the butterfly effect in chaos theory. Dubbed information scrambling (1–3), this quantum-mechanical effect gives rise to a variety of unusual pheno- mena and applications ranging from models of quantum gravity (4, 5) to quantum metrol- ogy (6). The speed of information scrambling can be quantified by out-of-time-ordered cor- relators (OTOCs) (7, 8), which constitute a measure of how fast the noncommutativity between two different quantum operations is established (9). In certain systems, the OTOC grows exponentially fast over time elQ t , where lQ > 0 defines the generalized quantum Lyapunov exponent (8). OTOCs have been measured (10) and used as probes for various many-body phenomena, such as ther- malization (11), quantum phase transitions (12), many-body entanglement growth (13), and quantum scrambling (14–16). However, the observation of exponential scrambling has remained elusive. One approach to effective time reversal in- volves changing the sign of the Hamiltonian ^H → (cid:1) ^H during the evolution of highly engi- neered quantum systems. In the field of quan- 1Department of Physics, MIT-Harvard Center for Ultracold Atoms, Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 2Department of Physics, Harvard University, Cambridge, MA 02138, USA. 3Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 4Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 5Viewpointsystem GmbH, 1010 Wien, Austria. *Corresponding author. Email: vuletic@mit.edu Fig. 1. Time-reversal–based exponential growth of sensitivity in a system with an unstable fixed point. (A) Classically, for a trajectory with a positive Lyapunov exponent l1 > 0, an initial signal (displacement) d 0ð Þ increases exponentially over time. For quantum dynamics, however, an initial overlap between two states is preserved under unitary evolution. To amplify the signal similarly to the classical case, one needs to evolve the ^ H, resulting in decreased quantum fluctuations along a direction with negative Lyapunov state under the nonlinear ^ coefficient l2 < 0. A displacement along this direction followed by application of the negative Hamiltonian (cid:1) H (such that l1;2→ (cid:1) l1;2) is then used to amplify the signal. The plots represent the evolution on the Bloch sphere, where the Sy and Sz axes are labeled as in the left subplot of the middle row. (B) Experimental setup. The LMG Hamiltonian is generated by the interaction of the collective atomic spin with light inside a cavity on the transition ↑j i→ ej i, while a radiofrequency magnetic field is applied to rotate the atomic spin. Li et al., Science 380, 1381–1384 (2023) 30 June 2023 1 of 4 RES EARCH | R E S E A R C H A R T I C L E unstable-fixed-point exponential evolution with a Lyapunov exponent lQ ¼ wj j. For a fixed Sc, choosing W ¼ Sc results in a maximum Lyapunov exponent lQ ¼ Scj j. Metrological gain Þ (cid:3)= S=2 ð ≡ maxa var Sað ½ We first measure the antisqueezing {largest variance x2 Þ} of the col- þ lective spin S after an evolution under ^H as a function of the ratio W= Scð Þ . The anti- squeezing x þconstitutes an upper bound on the quantum Fisher information (QFI) with respect to spin rotations (42). As shown in Fig. 2B, the experimental data for x2 þ agree with the numerical simulation of the model (solid red line) and show a peak at W ¼ Sc, as expected. We then measure in Fig. 2C how x2 þ grows with time for the two cases W ¼ 0 (OAT Hamil- tonian) and W ¼ Sc (critically tuned LMG Hamiltonian). The OAT data (gray) exhibit quadratic growth of x2 þ , as expected. The LMG data (red) show exponential growth of þ ¼ e2lQ t, with lQ ¼ Sc for times t ≲ ðScÞ(cid:1)1. x2 For larger times, the growths slows because of finite particle number and light-induced decoherence (34, 39). The finite total spin fur- ther causes the states to turn non-Gaussian, which we characterize via the Binder cumu- lant (43) (Fig. 2D). The time evolution under the critically tuned (W ¼ Sc) LMG Hamiltonian ^H quickly pre- pares an entangled collective quantum state. To implement quantum metrology with the SATIN protocol, we then apply a small rotation ^U df ¼ e(cid:1)i^S adf, where ^S a ≡ ^S ycosa þ ^S z sina represents a collective spin operator in the yz plane. This encodes a signal phase df along the a direction, with a ¼ p=4 chosen to max- imize the metrological gain [Fig. 1A and SM (39)]. To implement (cid:1) ^H, we switch to another set of laser frequencies incident on the cavity and flip the sign of the transverse field W [SM (39)]. This generates an effective backward evolution in time that amplifies the applied signal df. The shifted state then undergoes a bifurcated trajectory for df ≶ 0 (see Fig. 2C), resulting in an exponentially amplified devia- tion Gdf from the original position. We mea- sure the mean spin value h^S ai ¼ Ssin Gdf Þ to infer the signal amplification G. As shown in Fig. 3, the squared signal amplification G2 (orange) increases exponentially with the same exponent 2lQ as the antisqueezing x2 þ up to times t ≈ ðScÞ(cid:1)1. The measured quantum noise N 2 , i.e., the variance of spin projection noise along the amplification direction ^S a normalized to the standard quantum limit (SQL) (blue), remains unity until t ≈ 0:8ðScÞ(cid:1)1 . The subse- quent increase of the noise N 2 results from the residual light-atom entanglement (34) and can be improved in the future by optimizing the light detuning (39). The improvement of the metrological gain over the SQL is 6.8(4) dB. ð Fig. 2. Collective-spin evolution in the cQED system. (A) Numerical calculation of the normalized variance x2 þ of the antisqueezed direction as a function of Sct and W=Sc with linecuts representing the measurements in (B) and (C). (B) Experimentally measured antisqueezing (blue symbols) and theoretical expectation (red line) for Sct ¼ 1:8 as a function of the rotation strength W. The shaded region indicates exponential growth, whereas in the other regions the time evolution is either quasi-periodic or growing polynomially. (C) Comparison of antisqueezing x2 þ between the fastest exponential growth for a critical rotation strength W ¼ Sc, and the polynomial growth of pure OAT (W ¼ 0). The two Bloch spheres represent the lines of classical evolution in both situations. The dashed and dash-dotted red lines represent exponential growth based on the theoretically predicted Lyapunov exponent and the full numerical result, respectively, with no free parameters. The gray dashed line is calculated for W ¼ 0. (Inset) Logarithmic plot for W ¼ Sc showing the exponential growth of x2 þ. (D) The Binder cumulant B (characterizing the deviation from a Gaussian distribution for B > 0) for the antisqueezed direction for the critical LMG condition W ¼ Sc versus time t. (Insets) Measured spin distributions for Sct ¼ 0 (blue) and Sct ¼ 2 (purple), with the latter being strongly non-Gaussian. Fig. 3. Metrological gain with exponential LMG time-reversal protocol. The squared signal amplification G2 (pink open circles) and system noise N2 (blue solid squares) versus time t. The pink dashed line represents the exponen- tial growth of the antisqueezing shown in Fig. 2, and constitutes an upper bound to the QFI. The blue dash-dotted line is the calculated noise (with no free parameters) due to residual light- atom entanglement. The maximum metrological gain is 6:8 4ð Þ dB. The deviation of G2 from an exponential for t ≳ ðScÞ(cid:1)1 is due to the nonuniform coupling between atoms and the cavity light (44), as well as the residual light-atom entanglement, both of which can be improved in the future (34, 45). Quantum metrology and quantum information To investigate the quantum information sci- ence aspect of the time-reversal protocol, we measure the fidelity OTOC (FOTOC) with quan- tum state tomography using randomized mea- surements (39, 46, 47). The FOTOC F tð Þ can be expressed as the trace between the den- sity matrix r 0ð Þ of the original state and that of the state displaced by df evolved back- t 0ð Þ :¼ ^U t^r 0ð Þ ^Ut†, where ward in time, r′ ^U t :¼ ei ^Hte(cid:1)i^S adfe(cid:1)i ^Ht D F tð Þ ≡ ^U t^r 0ð Þ ^U † (cid:1) ¼ Tr r′ t 0ð Þr 0ð Þ t ^r 0ð Þ E (cid:3) At fixed forward evolution time t, the FOTOC F depends on the small displacement df and ð2Þ Li et al., Science 380, 1381–1384 (2023) 30 June 2023 2 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. FOTOC and OTOC extracted from quantum state tomography. (A) Experimental Wigner functions obtained from quantum state tomography after applying the LMG SATIN protocol with different signal displacements df [for W ¼ Sc and t ¼ 0:57ðScÞ(cid:1)1]. The dashed circle indicates the original CSS state. (B) The solid blue line is a quadratic fit to the data (open circles), used to extract the OTOC I (see text and Eq. 3). Fig. 5. Comparison between quantum information and quantum metrology parameters for the LMG model. The purple open circles, solid green squares, and solid blue diamonds represent the antisqueezing, metrological gain, and OTOC, respectively. All quantities increase initially exponentially with time with a fitted Lyapunov exponent lQ ¼ 1:01 3ð ÞSc that agrees well with the theoretical prediction lQ ¼ Sc. The OTOC error bars are obtained by using the bootstrapping method (39, 50). For longer times t ≳ ðScÞ(cid:1)1, the metrological gain and OTOC decrease owing to decoherence caused by light-atom entanglement, as is well captured by the theoretical model (solid blue line) without free parameters. is related to the OTOC I tð Þ by its second de- rivative (12) 1 2 @2F tð Þ I tð Þ ≡ (cid:1) (cid:5) ð@dfÞ2 jdf¼0 ¼ ^S a tð Þ^r 0ð Þ^S a tð Þ^r 0ð Þ (cid:6) ð3Þ with the Hermitian operator ^S a tð Þ :¼ ei ^Ht ^S ae(cid:1)i ^Ht. Choosing four different evolution times (such that Sct1 ∈ 0:38; 0:57; 0:77; 0:96 g), we f displace the entangled state for each t1 by several different small angles df. We then perform the tomographic reconstruction after a reversed time evolution with (cid:1) ^H to obtain F t1ð Þ, as shown in Fig. 4A. The OTOC I t1ð Þ is then extracted from the data by fitting a quad- ratic function in the displacement df to the FOTOC (Fig. 4B). We notice that the fitted quadratic curve is slightly shifted from df ¼ 0 and has slightly reduced peak fidelity. The shift is likely due to a small difference between the assumed and actual Larmor frequencies between the spin states, whereas the reduc- tion from unit peak fidelity is due to the im- perfect time reversal associated with residual light-atom entanglement. The small imperfec- tions do not reduce the metrological gain ap- preciably (39). Figure 5 summarizes our findings regard- ing the close relation between quantum scram- bling and time-reversal quantum metrology: The antisqueezing x2 þ, metrological gain G :¼ G2=N 2, and OTOC I all agree with each other and scale exponentially with application time t of the LMG Hamiltonian for t ≲ 0:8ðScÞ(cid:1)1. The exponential fit yields a Lyapunov exponent lQ= Scð Þ ¼ 1:01 T 0:03, in excellent agreement with the theoretical expectation lQ= Scð Þ ¼ 1. Concluding remarks Our experiments demonstrate a cQED real- ization of the critically tuned (W ¼ Sc) LMG model with an exponential evolution in phase space. We also point out and experimentally verify that time-reversal protocols represent a powerful experimental tool giving access not only to metrological gain beyond the SQL (22, 23, 48), but also enabling the measure- ment of quantum information scrambling in many-body systems. This generalizes the pos- sibilities of using quantum information science to enhance quantum metrology (49). Further- more, we observe exponential growth of both the OTOC and the metrological gain for the LMG model, thereby experimentally verify- ing the intrinsic relation between these two concepts from different subfields of quantum science. The demonstrated methods to re- verse time evolution may enable the exper- imental investigation of complex many-body quantum systems in which the information spreads exponentially fast within many de- grees of freedom. REFERENCES AND NOTES 1. Y. Sekino, L. Susskind, J. High Energy Phys. 2008, 065 (2008). 2. B. Swingle, G. Bentsen, M. Schleier-Smith, P. Hayden, Phys. Rev. A 94, 040302 (2016). 3. R. J. Lewis-Swan, A. Safavi-Naini, J. J. Bollinger, A. M. Rey, Nat. Commun. 10, 1581 (2019). 4. T. Schuster et al., Phys. Rev. X 12, 031013 (2022). 5. S. Nezami et al., PRX Quantum 4, 010321 (2023). 6. J. Appel et al., Proc. Natl. Acad. Sci. U.S.A. 106, 10960–10965 (2009). 7. A. Larkin, Y. N. Ovchinnikov, Sov. Phys. JETP 28, 1200 (1969). 8. A. Kitaev, A simple model of quantum holography [video] (2015). https://online.kitp.ucsb.edu/online/entangled15/ kitaev/. See also J. Maldacena, D. Stanford, Phys. Rev. D 94, 106002 (2016). I. García-Mata, R. A. Jalabert, D. A. Wisniacki, Scholarpedia 18, 55237 (2023). 9. 10. J. Li et al., Phys. Rev. X 7, 031011 (2017). 11. A. M. Green et al., Phys. Rev. Lett. 128, 140601 (2022). 12. K. X. Wei et al., Phys. Rev. Lett. 123, 090605 (2019). 13. M. Gärttner et al., Nat. Phys. 13, 781–786 (2017). 14. K. A. Landsman et al., Nature 567, 61–65 (2019). 15. J. Braumüller et al., Nat. Phys. 18, 172–178 (2022). 16. X. Mi et al., Science 374, 1479–1483 (2021). 17. E. Davis, G. Bentsen, M. Schleier-Smith, Phys. Rev. Lett. 116, 053601 (2016). 18. F. Fröwis, P. Sekatski, W. Dür, Phys. Rev. Lett. 116, 090801 (2016). 19. O. Hosten, R. Krishnakumar, N. J. Engelsen, M. A. Kasevich, Science 352, 1552–1555 (2016). 20. S. P. Nolan, S. S. Szigeti, S. A. Haine, Phys. Rev. Lett. 119, 193601 (2017). 21. T. Macrì, A. Smerzi, L. Pezzè, Phys. Rev. A 94, 010102 (2016). 22. K. A. Gilmore et al., Science 373, 673–678 (2021). 23. S. Colombo et al., Nat. Phys. 18, 925–930 (2022). 24. T.-W. Mao et al., Quantum enhanced sensing by echoing spin-nematic squeezing in atomic Bose-Einstein condensate (2022). arXiv:2212.09124. 25. H. J. Lipkin, N. Meshkov, A. Glick, Nucl. Phys. 62, 188–198 (1965). Li et al., Science 380, 1381–1384 (2023) 30 June 2023 3 of 4 RES EARCH | R E S E A R C H A R T I C L E 26. L.-M. Duan, A. Sørensen, J. I. Cirac, P. Zoller, Phys. Rev. Lett. 43. K. Binder, Z. Phys. B Condens. Matter Quanta 43, 119–140 85, 3991–3994 (2000). (1981). 27. T. Zibold, E. Nicklas, C. Gross, M. K. Oberthaler, Phys. Rev. Lett. 44. J. Hu, W. Chen, Z. Vendeiro, H. Zhang, V. Vuletić, Phys. Rev. A 105, 204101 (2010). 92, 063816 (2015). 28. C. D. Hamley, C. Gerving, T. Hoang, E. Bookjans, M. S. Chapman, Nat. Phys. 8, 305–308 (2012). 29. H. Strobel et al., Science 345, 424–427 (2014). 30. W. Muessel et al., Phys. Rev. A 92, 023603 (2015). 31. J. Peise et al., Nat. Commun. 6, 8984 (2015). 32. S. Pilatowsky-Cameo et al., Phys. Rev. E 101, 010202 (2020). 33. J. A. Muniz et al., Nature 580, 602–607 (2020). 34. Z. Li et al., PRX Quantum 3, 020308 (2022). 35. M. Kitagawa, M. Ueda, Phys. Rev. A 47, 5138–5143 (1993). 36. S. Pappalardi et al., Phys. Rev. B 98, 134303 (2018). 37. E. B. Rozenbaum, L. A. Bunimovich, V. Galitski, Phys. Rev. Lett. 125, 014101 (2020). 38. H. Tanji-Suzuki et al., in Advances in Atomic, Molecular, and Optical Physics, E. Arimondo, P. R. Bermna, C. C. Lin, Eds. (Academic Press, 2011), vol. 60, pp. 201–237. 39. See supplementary materials for details. 40. T. Holstein, H. Primakoff, Phys. Rev. 58, 1098–1113 (1940). 41. C. K. Law, H. T. Ng, P. T. Leung, Phys. Rev. A 63, 055601 (2001). 42. L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, P. Treutlein, Rev. Mod. Phys. 90, 035005 (2018). 45. B. Wu, G. P. Greve, C. Luo, J. K. Thompson, Site-dependent selection of atoms for homogeneous atom-cavity coupling. arXiv:2104.01201 [quant-ph] (2021). 46. R. Schmied, P. Treutlein, New J. Phys. 13, 065019 (2011). 47. A. Elben et al., Nat. Rev. Phys. 5, 9–24 (2023). 48. D. Barberena, S. R. Muleady, J. J. Bollinger, R. J. Lewis-Swan, A. M. Rey, Fast generation of spin squeezing via resonant spin- boson coupling. arXiv:2201.10622 [quant-ph] (2022). 49. C. D. Marciniak et al., Nature 603, 604–609 (2022). 50. W. Hardle, J. S. Marron, Ann. Stat. 19, 778 (1991). 51. Z. Li et al., Improving metrology with quantum scrambling, Dryad, dataset (2023). https://doi.org/10.5061/dryad.zkh1893fj. ACKN OWLED GMEN TS We thank J. Thompson, M. Schleier-Smith, B. Braverman, and A. Adiyatullin for inspiring discussions. We also thank M. Liu for helping resolve the subtleties of using bootstrap method to extract statistical uncertainties. Funding: We acknowledge financial support from ONR (grant N00014-20-1-2428), the National Science Foundation through the Center for Ultracold Atoms (grant PHY-1734011) and NSF QLCI (Award OMA - 2016244), DOE QSA (through grant DE-SC0021013 and LBNL QSA Center), and ARO (grant W911NF1910517). Author contributions: Z.L., S. Colombo, C.S., G.V., and E.P. contributed to building the experimental setup and performed the measurements. Z.L. and S.Colombo analyzed the data, numerically simulated the experiments, and contributed to theoretical interpretation of the results. R.S. contributed to developing the quantum state tomography method. S.P.-C. and S. Choi contributed to the theoretical interpretation of the results. All work was supervised by V.V. All authors discussed the results and contributed to the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the figures of the paper and/or the supplementary materials. The data can be accessed at Dryad (51). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg9500 Materials and Methods Figs. S1 and S2 Tables S1 and S2 References (52–58) Submitted 1 February 2023; accepted 19 May 2023 10.1126/science.adg9500 Li et al., Science 380, 1381–1384 (2023) 30 June 2023 4 of 4
10.1126_science.adh1411
RES EARCH STRUCTURAL BIOLOGY Translation dynamics in human cells visualized at high resolution reveal cancer drug action Huaipeng Xing1,2, Reiya Taniguchi1, Iskander Khusainov1, Jan Philipp Kreysing1,3, Sonja Welsch4, Beata Turonˇová1, Martin Beck1,5* Ribosomes catalyze protein synthesis by cycling through various functional states. These states have been extensively characterized in vitro, but their distribution in actively translating human cells remains elusive. We used a cryo–electron tomography-based approach and resolved ribosome structures inside human cells with high resolution. These structures revealed the distribution of functional states of the elongation cycle, a Z transfer RNA binding site, and the dynamics of ribosome expansion segments. Ribosome structures from cells treated with Homoharringtonine, a drug used against chronic myeloid leukemia, revealed how translation dynamics were altered in situ and resolve the small molecules within the active site of the ribosome. Thus, structural dynamics and drug effects can be assessed at high resolution within human cells. tive positions. By contrast, the elongation factor bound “eEF1A, A/T, P” state was most abundant in eukaryotic D. discoideum cells. In microsomes isolated from human cells, the membrane-associated elongation cycle has been resolved, revealing the eEF1A, A/T, P, E state as the most prominent (8). However, to which extent these insights are applicable to actively translating human cells remains unclear. A eEF1A, A/T, P B Homoharringtonine (HHT) is a natural compound that binds to the ribosome and in- hibits protein synthesis (9). It is used to treat patients with chronic myeloid leukemia clini- cally and as a reference to study new anticancer ribosome inhibitors (9, 10). The HHT-bound ribosome structure has been determined by incubating the purified 80S ribosome with HHT in vitro (11, 12), revealing the binding site at the peptidyl transferase center (PTC). In the cellular context, it remains unclear how HHT affects translation dynamics and how exactly it inhibits protein synthesis. We applied cryo– focused ion beam (cryo-FIB), cryo-ET and advanced data processing algorithms to de- termine the near-atomic structures of ribosomes and analyzed the abundance and organization of different ribosome states in untreated and HHT-treated human cells. Ribosomes are bound with HHT inside human cells To study the 80S ribosome structures inside human cells, we first prepared cryo–FIB-milled lamellae from 35 native (untreated) human embryonic kidney cells (HEK-293) and acquired 358 tilt series (fig. S1A and table S1). We used template matching to identify ribosomes within the reconstructed tomograms (fig. S1B and T he eukaryotic ribosome (80S) consists of two subunits (60S and 40S) that trans- late mRNA into proteins (1). Purified 80S ribosomes have been extensively studied in vitro, which has elucidated molecular details of translation (2, 3). The translation process can be divided into four main stages: initiation, elongation, termination, and ribo- some recycling (3). During elongation, the rotation of the 40S, the association of elonga- tion factors, and the translocation of tRNAs are coordinated to synthesize nascent chains (3). tRNAs have three canonical binding sites on the ribosome: aminoacyl (A), peptidyl (P), and exit (E) sites (3). A noncanonical tRNA binding site called the Z site, located in an extreme position past the E site, has been identified in structures of isolated transla- tionally inactive mammalian ribosomes (4). It is speculated that the Z site may represent a late-stage intermediate of tRNA ejection downstream of the E site with similarity to internal ribosome entry site interactions, but the exact physiological relevance remains ob- scure (4). The 70S ribosome from Mycoplasma pneu- moniae and 80S ribosome from Dictyostelium discoideum and the respective translation elongation processes have been visualized inside cells using cryo–electron tomography (cryo-ET), revealing differences in translation elongation (5–7). For example, the most abun- dant elongation intermediate in bacterial M. pneumoniae cells was the “A, P” state wherein tRNAs were identified in the respec- 1Department of Molecular Sociology, Max Planck Institute of Biophysics, 60438 Frankfurt am Main, Germany. 2Faculty of Biochemistry, Chemistry and Pharmacy, Goethe University Frankfurt am Main, 60438 Frankfurt am Main, Germany. 3IMPRS on Cellular Biophysics, 60438 Frankfurt am Main, Germany. 4Central Electron Microscopy Facility, Max Planck Institute of Biophysics, 60438 Frankfurt am Main, Germany. 5Institute of Biochemistry, Goethe University Frankfurt, Frankfurt am Main, Germany. *Corresponding author. Email: martin.beck@biophys.mpg.de Xing et al., Science 381, 70–75 (2023) 7 July 2023 40S 60S P A/T Codon sampling Codon recognition eEF1A 2.7 4.0 5.0 6.0 7.0 8.0 C D Sampling P site A site mRNA A 1824 A 1825 18S rRNA Fig. 1. 80S ribosome structures in human cells. (A) Color-coded local resolution map of the eEF1A, A/T, P state identified in the untreated dataset. (B) Superposition of the codon recognition state (PDB 5LZS, cyan) onto our eEF1A-tRNA ternary complex at the eEF1A, A/T, P state. The atomic model of tRNA (lavender) and eEF1A (maroon) from PDB 4CXG were fitted into our map for comparison to 5LZS. (C) The mRNA segmented from the eEF1A, A/T, P state was fitted with PDB 5LZS. Density is more clearly defined at the P site as compared with the A site. (D) The atomic model (5LZS) of 18S rRNA (A1822 to A1827) was rigid-body–fitted into the corresponding density of the 80S at the eEF1A, A/T, P state. The decoding nucleotides A1824 and A1825 in the atomic model are in the flipped-out configuration, which is inconsistent with the observed density (arrowhead). 1 of 6 RES EARCH | R E S E A R C H A R T I C L E movie S1). After classification and refinement, we determined the ribosome structure at an overall resolution of ~3.2 Å and locally resolved features up to 2.5 Å in resolution from un- treated cells (Fig. 1B and figs. S1 to S3). The den- sity of ribosomal protein side chains, ribosomal RNA (rRNA) bases, tRNA bases, and ions were well-resolved (fig. S1E), indicating the quality of the map. Using the same approach we processed 352 tilt series from 32 HHT-treated cells. In the resulting structure the density of HHT was visi- ble (fig. S1, F and G) and the position of HHT at the PTC showed a high similarity to previous in vitro analysis (9, 11). The P-tRNA was poorly resolved compared with the untreated ribosome (fig. S1G), suggesting that the translation was altered after HHT treatment. Translation elongation cycle in human cells We next investigated the translation elonga- tion process. In untreated cells, focused classi- fication was performed with dedicated tRNA and elongation factor masks, resulting in eight ribosome states resolved from 3.4 Å to 16.4 Å (figs. S2 and S3 and table S2). These states were well-explained by previously reported structural models that captured various elon- gation states (figs. S4 to S6) (4, 8, 13–15). The eEF1A, A/T, P state was the best resolved, globally reaching ~3.4 Å (Fig. 1A), and dis- played structural features characteristic for tRNA scanning, indicating that codon sampling A 60S eEF1A-tRNA rather than codon recognition is occurring (Fig. 1B and fig. S4A) (13, 16). Specifically, the mRNA codon was more clearly resolved at the P site (Fig. 1C), suggesting that the codon- anticodon interaction at the A site is not yet established. Consistently, we observe the decod- ing nucleotides A1824 and A1825, which are thought to stabilize the A-site tRNA by a “flipping out” movement upon codon recogni- tion (13), in a “flipped in” conformation (Fig. 1D). In total, six classes—accounting for ~83% of all identified ribosomes—were assigned to the translation elongation cycle based on their similarity to the previously reported structures (Fig. 2A) (3, 5, 13, 16). The elongation cycle is thought to start from the non-rotated P state that was detected but not particularly abun- dant in our data (5) (Fig. 2A). It is followed by the eEF1A, A/T, P state discussed above which was the most prominent, consistent with in situ analysis of the lower eukaryote Dictyo- stelium discoideum (7), but in contrast to pre- vious analysis in bacteria (5). The abundance of the codon sampling state in human cells could be important for the higher decoding fidelity during translation (17). The following states were the non-rotated A/T, P and A, P states. At a low contour level of the map a factor- like density at the eEF1A binding region is seen in the A/T, P state, but neither an extended nor compact eEF1A model provides a sufficient explanation (fig. S4C, Materials and Methods) (8, 13). As expected, the A, P state displayed the A1824 and A1825 nucleotides in a flipped out conformation (fig. S4D). Yet it was less abun- dant than in bacteria, where it was the most populated (5). The subsequent state was iden- tified as eEF2, ap/P, pe/E (18). Next, two possible sequences of events are conceivable: eEF2 and E-tRNA could first dissociate from the ribosome forming the P state that subse- quently binds the eEF1A-tRNA, or the eEF1A- tRNA may displace eEF2 with subsequent departure of the E-tRNA (5) (Fig. 2A). Various conformational changes of the small subunit are apparent during the elongation cycle (fig. S5) and consistent with previous work (16, 18). Two further states were identified whose places in the elongation cycle are not obvious (Fig. 2B). The rotated eEF2 state without tRNA fitted the previous structural model very well (15, 19) (fig. S4G), although the eEF2 position was shifted by ~1.7 nm compared with the eEF2, ap/P, pe/E state (fig. S4I). The remain- ing class, which accounts for only 1.6% of all classified ribosomes (Fig. 2B), was consistent with the non-rotated eEF1A, A/T, P, Z state, which has been suggested to be transient (4). Although the resolution of this state was only moderate in untreated cells, the observed elec- tron optical density at the Z site fitted well with the purified ribosome at the non-rotated P, Z state (Fig. 2B and fig. S4, H and J) (4). To our knowledge, alternative factors binding to this B R2 eEF2 3.8% C 60S Untreated HHT-treated N eEF1A, A/T, P, Z 1.6% eEF1A-tRNA Compact eEF2 40S N P 10.2% eEF1A R1+R2 A, P 15.3% R1+R2 Compact eEF2, A, P 3.4% 40S N P 4.5% eEF1A-tRNA N eEF1A, A/T, P 39.6% R2 eEF2, ap/P, pe/E 5.2% eEF2 N eEF1A, A/T, P, E 13.0% R1+R2 A, P 14.1% N A/T, P 6.4% D R1: Rolled R2: Rotated N: Non-rotated and unrolled R2 eEF2 46.9% N eEF1A, A/T, P, Z 6.7% N A/T, P, Z 3.3% Fig. 2. Ribosome states in native untreated and HHT-treated cells. (A) Six ribosome states are assigned to the elongation cycle in untreated cells. (B) The eEF2 and eEF1A, A/T, P, Z states in untreated cells. (C) Three potential translation elongation intermediate states in the HHT-treated cells. The dashed arrow illustrates how the elongation states may connect. (D) The eEF2, eEF1A, A/T, P, Z and A/T, P, Z states inside the treated cells. R1, rolled; R2, rotated; N, nonrotated and unrolled. The information about 40S rolling and rotation is shown in figs. S5 and S11. Xing et al., Science 381, 70–75 (2023) 7 July 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E site have not been identified. More importantly, we obtained higher resolution in HHT-treated cells (see below), which further strengthened this conclusion. We note that ~12% of the particles identified as ribosomes were not at- tributed to a specific state by our classification and may correspond to additional low abun- dant states or transitions between states. Col- lectively, our data allowed the detailed analysis of the translation dynamics inside native hu- man cells, which exhibited differences compared with previous ex vivo studies (8, 18), indicating that the ribosome isolation process may affect the ability to capture native elongation inter- mediates (5, 7). HHT alters the elongation cycle The anticancer drug HHT blocks peptide bond formation (9, 20). Previous in vitro analy- sis of purified ribosomes revealed that HHT binds in the A-site cleft in the PTC (11, 21). Upon HHT treatment, one would thus expect to observe a strong enrichment of ribosomes with the A- and P-site tRNAs bound (11). Indeed, M. pneumoniae ribosomes treated with 0.5 mg/ml (1547.4 mM) chloramphenicol (Cm), a related antibiotic binding to the PTC of the 50S, showed over 70% of ribosomes at the A, P state (5, 22). The effect of HHT on human ribosome states in cells remains unknown. To address this, we exposed cells to HHT concen- trations of 0.055 mg/ml (100 mM) in the medium, which resulted in a significantly lower cellu- lar protein concentration than untreated cells (fig. S7A), implying that protein synthesis was inhibited. At the time point examined by cryo-EM, the ATP levels, which serve as an indicator of cell viability, were not yet reduced (fig. S7B) and the cell morphology of untreated and treated cells was indistinguishable (fig. S7C). Classification, as described above, identified six ribosome states resolved into the 3.7 Å to 11.5 Å range (figs. S8 to S12 and table S3). Unexpectedly, the rotated state with eEF2 but without tRNA accounted for almost half of all ribosomes in the dataset (Fig. 2, C and D, and Fig. 3A). This state showed density for SERPINE1 mRNA-binding protein 1 (SERBP1) and thus accounts for hibernating ribosomes (15, 23) (fig. S10G), in which the HHT density was visible (fig. S8C). The P state increased in abundance compared with the untreated cells, whereby the A, P states showed a similar abun- dance (Fig. 2C). The HHT molecule was also resolved at the A, P state but not in other less populated classes as a result of the lower resolution (Fig. 3B). We therefore combined the particles from the four remaining less abun- dant classes. The resulting average also displayed density for HHT (fig. S8C). The positions of A-tRNA and P-tRNA at A, P states were simi- lar in untreated and HHT-treated cells (Fig. 3B) (9). One of our classes appeared similar to a potential compact eEF2, A, P state (figs. S10, D, I, and J, and S11F) (24), although the respective local resolution prevents a definitive assign- ment due to a lack of secondary structure. Finally, two states with Z-tRNA were consider- ably more abundant compared with untreated cells (Fig. 2D and fig. S10, C and E), showing the typical features of tRNA shape at the Z site (fig. S10, E and H, and movie S3). Thus, HHT treatment results in the accumula- tion of ribosome hibernation instead of the A, P state, which may be representative of the mechanism of the drug action in cancer therapy. To investigate cell-to-cell variability, we ana- lyzed the distribution of the ribosome states across individual cells captured by multiple A e g a t n e c r e P e g a t n e c r e P 50 40 30 20 10 0 50 40 30 20 10 0 B A, P A, P Untreated A 4411 G 3878 G 4413 U 4414 A 4449 G 4451 G 3907 U 4452 HHTHHT 90° 1 2 3 4 5 6 7 8 9 10 Treated 1 2 3 4 5 6 7 8 9 10 Ribosome class 1 eEF1A, A/T, P 4 eEF1A, A/T, P, E 8 eEF1A, A/T, P, Z 2 A/T, P 5 eEF2 9 A/T, P, Z 3 eEF2, ap/P, pe/E 6 P 10 Compact eEF2, A, P 7 A, P E P A HHT Fig. 3. Distribution of ribosome states in cells. (A) Percentages of ribosome states in untreated and HHT-treated cells. (B) The PTC of ribosomes at the A, P state in untreated and treated cells is shown fitted with PDBs 5AJ0 and 6QZP, respectively. Structural overlay of tRNAs at A, P states with models determined from previous studies (bottom; materials and methods). The black arrow points to P-tRNA. HHT is colored in cyan. tomograms. We identified some degree of cell- to-cell variability (fig. S13A), for example, the eEF1A, A/T, P state varied from 25 to 56% in untreated cells (fig. S13B). Overall, the signal observed in multiple tomograms of the same cell was similar, with some notable exceptions that may imply local variability (fig. S13B). Notably, the abundance of the eEF1A, A/T, P (class 1) and eEF1A, A/T, P, E (class 4) was largely anticorrelated in untreated cells (fig. S13, C and D), whereas the sum of eEF1A-bound states was relatively consistent, suggesting that its total concentration may be limited (fig. S13E). By contrast, the abundance of ribosomes containing eEF2 was much more diverse be- tween different cells (fig. S13E). Polysomes are impaired upon HHT treatment We analyzed the spatial organization of dif- ferent ribosome states within polysomes in treated and untreated cells. The densities ac- counting for neighboring ribosomes that are commonly observed in the subtomogram aver- ages of untreated cells were reduced in HHT- treated cells (fig. S14A). Furthermore, the ribosome distribution was more dispersed sub- sequent to treatment (Fig. 4, A and B), both implying that polysomes might be largely abol- ished. To test this hypothesis, we defined poly- somes based on a distance threshold of 9 nm from the mRNA exit to entry sites of neighbor- ing particles (fig. S14B and table S4, Materials and Methods). Although this arbitrary classifier will not capture the entirety of polyribosomes as it could, for example, miss more distantly spaced ribosomes on a longer transcript, over- all it performs well in their detection (fig. S15). Of all ribosomes in untreated cells, 30.2% were grouped into arbitrary polysomes (Fig. 4C and fig. S14, C and D), which was considerably higher than in HHT-treated cells (fig. S15, A to E). In monosomes the eEF2 state was prevalent (Fig. 4D and fig. S16, A and B) and exhibits much less neighboring density (fig. S16, C and D), under- lining the notion that these ribosomes are hi- bernating. In polysomes, the eEF2, ap/P, pe/E state (class 3, see Fig. 3A) was less frequent in the first leading ribosome than the trailing ribo- somes (fig. S15F). Classification of all untreated ribosomes with a dedicated mask resolved the low-resolution structure of the di-ribosome, in which the mRNA density was visible (fig. S17A) (25, 26). This di-ribosome resembled the top- to-top configuration (t-t, the central protuber- ance of both ribosomes facing a similar direction) (27, 28) (fig. S17B). Two other arrangements of pairs were apparent: one with the central protuberance of the i+1 ribosome toward down (t-d), the other toward up (t-u) (fig. S17, C and D). Although the abundance of these configurations declined in the treated cells (fig. S17, D and E), the center-to-center and exit-to-entry distance of these pairs were indistinguishable from un- treated cells (fig. S17F). Xing et al., Science 381, 70–75 (2023) 7 July 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E A Untreated B HHT-treated 1 2 3 4 5 6 7 C 90° 3’ 5’ 3’ 5’ 8 D e g a t n e c r e P 30 25 20 15 10 5 0 1 2 5 6 7 8 9 10 e g a t n e c r e P 50 40 30 20 10 0 5 7 8 Monosomes Polysomes 9 10 8 7 6 Ribosome class HHT-treated 6 5 4 3 Ribosome class Untreated Fig. 4. Spatial and functional analysis of polysomes. (A and B) Distribution of ribosome states in a representative tomogram from an untreated (A) and treated (B) cell. (C) A polysome and the putative mRNA path. The center-to-center distance of the neighboring ribosomes: 27.3 ± 1.8 nm (mean ± SD, n = 8). (D) Distribution of ribosome states in monosomes and polysomes from 358 tomograms of untreated cells and 352 tomograms of treated cells. Class numbers are the same as in Fig. 3A. We classified ribosomes with a mask cover- ing the region beyond the exit tunnel, thus capturing both the potential membrane and the expansion segment ES27L (figs. S18 and S19, and tables S2 and S3). The analysis re- vealed that the translation state distributions of soluble and membrane-bound ribosomes were similar (fig. S18D). Notably, seven con- formations of the expansion segment ES27L were found in cytosolic ribosomes (fig. S19, Materials and Methods). One of these con- tains a long stretch of ES27L that associates with the ErbB3 receptor-binding protein (Ebp1) on the surface of the 60S and resembles a previously published in vitro structure (29, 30) (fig. S19, A and B, and tables S2 and S3). The percentage of Ebp1-associated ribosomes was below 20% in most untreated cells, whereas the abundance was over 25% in all HHT-treated cells (fig. S20A). Whether Ebp1 is related to spe- cific translation states remained unclear (29, 30). Xing et al., Science 381, 70–75 (2023) 7 July 2023 Thus, a subfraction of Ebp1-bound ribosomes is observed in all apparent intermediates of the elongation cycle (fig. S20B). HHT binds the free 60S associated with eIF6 Finally, we investigated the free 60S and 40S in the cytoplasm of human cells. The free 60S showed Ebp1 and ES27L density in both data- sets (Fig. 5, A and B, fig. S21, A and B), which was similar to the Ebp1-associated 80S ribo- some. However, the 60S was much more abun- dant in treated cells than in untreated cells (Fig. 5C). Eukaryotic initiation factor 6 (eIF6) binds the 60S to prevent premature 60S bind- ing with 40S (31, 32). Our data show that the 60S did not contain eIF6 in the cytosol of native untreated cells, contrasting HHT-treated cells (Fig. 5D). Thus, it appears that eIF6 prevents the association between large and small ribosomal subunits in cells (33–35) (Fig. 5, C and D, and fig. S21, C to E). Furthermore, the HHT density was resolved in the 60S from treated cells (Fig. 5E). Thus, we conclude that HHT can bind not only the PTC of different states of the 80S ribosome (Fig. 3B and fig. S8C) but also the free 60S decorated with eIF6, which may block the assembly of 40S and 60S in the cytosol of the cell. Discussion Our extensive cryo-ET analysis stresses the feasibility of obtaining structures at high res- olution and visualizes an anticancer drug inside human cells. The local resolution in the ribosome core in our study is limited by the pixel size. This illustrates that technical obsta- cles that hinder high resolution in cellular to- mography have been overcome. The combined improvements in energy filtering (36), image processing (Warp-RELION-M pipeline) (22) and tight control over specimen thickness (37, 38) enable high resolution, as long as the 4 of 6 ) % ( e c n a d n u b a S 0 6 20 15 10 15.5 5 4.1 0 Untreated Treated RES EARCH | R E S E A R C H A R T I C L E A D E B C 150° 40° 150° 40° 10.0 Å 90° ES27L Ebp1 4.2 Å Ebp1 90° Untreated Treated eIF6 eIF6 A 4449 A 4449 G 4451 G 3907 G 4451 G 3907 U 4452 HHT U 4452 HHT 90° HHT Fig. 5. Structure of 60S in human cells. (A and B) Structures of free 60S in the cytoplasm of untreated cells (A) and HHT-treated cells (B). ES27L and Ebp1 (PDB: 6SXO) are fitted into the 60S maps. Although Ebp1 was not confidently assigned to the 60S from untreated cells as a result of the lower resolution, it fitted better in comparison to alternative factors (amino-terminal acetyltrans- ferases and nascent polypeptide-associated complex) binding to the tunnel exit. (C) Percentage of the 60S in untreated and HHT-treated cells normalized to the number of 80S ribosomes in the respective dataset. In untreated cells, abundance = 60S/(60S+80S) = 1,693/(1693+39,402). In treated cells, 60S/ (60S+80S) = 7176/(7176+39,070). (D) The structure of 60S in untreated cells fitted with PDB 6LSR indicates that eIF6 is missing (left), contrasting HHT- treated 60S (right). Large subunit GTPase 1 (LSG1), NMD3 and ZNF622 (PDB: 6LSR) are not observed in the HHT-treated 60S structure. (E) The PTC of free 60S from untreated and treated datasets fitted with PDB 6QZP. HHT is colored in cyan. number of particles that can be obtained is sufficient. Further improvements in the autom- atization of specimen preparation techniques may thus be instrumental in pushing the reso- lution for other macromolecular assemblies analyzed inside cells (38–43). The detailed comparison of ribosome struc- tures within untreated and HHT-treated cells revealed that HHT binds to the PTC of the 80S ribosome in situ (Fig. 3B and figs. S1G and S8C), where it can block peptide bond forma- tion as suggested by previous in vitro studies (9, 11, 21). However, we also observed SERBP1 binding to ribosomes, a factor that functions downstream of mammalian target of rapamycin complex 1 (mTORC1) and leads to a dormant ribosome state (44). Such a dormant ribosome state can be a result of cellular stress signaling (44). Further, we uncovered that HHT bound the free 60S associated with eIF6 (Fig. 5, D and E), thus likely hampering the assembly of func- tional 80S ribosomes and in turn further impair- ing overall protein synthesis. The accumulation of 60S-eIF6 may be explained by the fact that the HHT binding site overlaps with that of the N terminus of Shwachman-Bodian-Diamond syndrome protein (SBDS) (fig. S21F) and that the binding of SBDS to 60S is necessary to release eIF6 (34, 45). Our study provides insights into the trans- lation elongation cycle, the coordination of ribosome activity within polysomes, and the diverse arrangements of ES27L inside human cells. The same technology may be used in the future to investigate ribosome, translation, and mRNA quality control pathways in the context of human tissue culture cells and primary patient-derived cells. Intermediates that have not yet been observed inside cells may be enriched by introducing kinetical bottlenecks through genetic or pharmaceu- tical perturbation (5, 46, 47). Our study may set the stage for the analysis of structures inside mammalian cells and allow characteri- zation of drug susceptibility of human indi- viduals at high resolution by cryo-ET. REFERENCES AND NOTES 1. V. Ramakrishnan, Cell 159, 979–984 (2014). 2. A. A. Korostelev, Annu. Rev. Biochem. 91, 245–267 (2022). 3. A. P. Schuller, R. Green, Nat. Rev. Mol. Cell Biol. 19, 526–541 (2018). 4. A. Brown, M. R. Baird, M. C. J. Yip, J. Murray, S. Shao, eLife 7, e40486 (2018). 5. L. Xue et al., Nature 610, 205–211 (2022). 6. F. J. O’Reilly et al., Science 369, 554–557 (2020). 7. P. C. Hoffmann et al., Nat. Commun. 13, 7435 (2022). 8. M. Gemmer et al., Nature 614, 160–167 (2023). 9. A. Gilles et al., Cells 9, 629 (2020). Xing et al., Science 381, 70–75 (2023) 7 July 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 10. V. Kumar, M. Jyotirmayee, Mol. Cell. Biochem. 478, 1–17 (2023). 11. N. Garreau De Loubresse et al., Nature 513, 517–522 (2014). 12. S. K. Natchiar, A. G. Myasnikov, H. Kratzat, I. Hazemann, B. P. Klaholz, Nature 551, 472–477 (2017). 13. S. Shao et al., Cell 167, 1229–1240.e15 (2016). 14. R. Buschauer et al., Science 368, eaay6912 (2020). 15. J. N. Wells et al., PLOS Biol. 18, e3000780 (2020). 16. T. V. Budkevich et al., Cell 158, 121–131 (2014). 17. A. Rozov, N. Demeshkina, E. Westhof, M. Yusupov, G. Yusupova, Trends Biochem. Sci. 41, 798–814 (2016). 18. E. Behrmann et al., Cell 161, 845–857 (2015). 19. P. R. Smith et al., Nat. Commun. 12, 6789 (2021). 20. S. Pellegrino, S. Terrosu, G. Yusupova, M. Yusupov, Cancers 13, 4392 (2021). 21. G. Gürel, G. Blaha, P. B. Moore, T. A. Steitz, J. Mol. Biol. 389, 146–156 (2009). 22. D. Tegunov, L. Xue, C. Dienemann, P. Cramer, J. Mahamid, Nat. Methods 18, 186–193 (2021). 23. P. R. Smith, S. C. Pandit, S. Loerch, Z. T. Campbell, Trends Biochem. Sci. 47, 477–491 (2022). 24. J. Lin, M. G. Gagnon, D. Bulkley, T. A. Steitz, Cell 160, 219–227 (2015). 25. K. Ikeuchi et al., EMBO J. 38, e100276 (2019). 26. M. Narita et al., Nat. Commun. 13, 6411 (2022). 27. F. Brandt et al., Cell 136, 261–271 (2009). 28. F. Brandt, L. A. Carlson, F. U. Hartl, W. Baumeister, K. Grünewald, Mol. Cell 39, 560–569 (2010). 29. K. Wild et al., Nat. Commun. 11, 776 (2020). 30. M. L. Kraushar et al., Mol. Cell 81, 304–322.e16 (2021). 31. D. W. Russell, L. L. Spremulli, J. Biol. Chem. 254, 8796–8800 (1979). 32. M. Gartmann et al., J. Biol. Chem. 285, 14848–14851 (2010). 33. M. Ceci et al., Nature 426, 579–584 (2003). 34. F. Weis et al., Nat. Struct. Mol. Biol. 22, 914–919 (2015). 35. P. Jaako et al., Nat. Commun. 13, 1562 (2022). 36. M. Obr et al., J. Struct. Biol. 214, 107852 (2022). 37. A. Rigort, J. M. Plitzko, Arch. Biochem. Biophys. 581, 122–130 (2015). 38. S. Klumpe et al., eLife 10, e70506 (2021). 39. M. Smeets et al., Micros. Today 29, 20–25 (2021). 40. S. Klumpe et al., Micros. Today 30, 42–47 (2022). 41. Z. Wang et al., Science 375, eabn1934 (2022). 42. H. Van den Hoek et al., Science 377, 543–548 (2022). 43. S. Mosalaganti et al., Science 376, eabm9506 (2022). 44. S. Shetty, J. Hofstetter, S. Battaglioni, D. Ritz, M. N. Hall, EMBO J. 42, e112344 (2023). 45. P. Jaako et al., Nat. Commun. 13, 1562 (2022). 46. M. Thoms et al., Science 369, 1249–1256 (2020). 47. S. Klinge, J. L. Woolford, Nat. Rev. Mol. Cell Biol. 20, 116–131 (2018). 48. turonova/cryoCAT: Initial release, Version v0.1.0, Zenodo (2023); https://zenodo.org/record/7997724. ACKN OWLED GMEN TS We thank P. C. Hoffmann, M. Tuijtel, and T. Majtner from the Department of Molecular Sociology at the Max Planck Institute of Biophysics for advice on membrane segmentation, data collection and data processing. We thank all the members from the Central Electron Microscopy facility and the Department of Molecular Sociology at the Max Planck Institute of Biophysics, and A. Schwarz, E. Schuman and the Department of Synaptic Plasticity at the Max Planck Institute of Brain Research for support and advice. We acknowledge the support from the Max Planck Computing and Data Facility. We thank M. Rodnina, N. Fischer, and S. Boehm for critical reading of the manuscript. Funding: M.B. acknowledges funding from the Max Planck Society. Author contributions: Conceptualization: H.X. and M.B. Methodology: H.X., B.T., I.K., and J.P.K. Investigation: H.X., R.T., I.K., J.P.K., S.W., and B.T. Visualization: H.X., B.T., and M.B. Funding acquisition: M.B. Project administration: H.X. and M.B. Supervision: H.X. and M.B. Writing – original draft: H.X., B.T., and M.B. Writing – review and editing: H.X., R.T., I.K., J.P.K., S.W., B.T., and M.B. Competing interests: The authors declare no competing interests. Data and materials availability: The original tilt series have been deposited in the Electron Microscopy Public Image Archive (EMPIAR-11538). EM maps have been deposited in the Electron Microscopy Data Bank (EMDB) under accession numbers: EMD-16721, 16725, 16726, 16727, 16728, 16733, 16734, 16735, 16736, 16737, 16738, 16739, 16740, 16741, 16742, 16743, 16722, 16744, 16748, 16747, 16749, 16750, 16751, 16752, 16754, 16755, 16756, 16757. The codes used for polysome analysis are part of the cryoCAT repository (48) and are deposited at Zenodo. All other data are available in the manuscript or supplementary materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh1411 Materials and Methods Figs. S1 to S21 Tables S1 to S4 References (49–64) MDAR Reproducibility Checklist Movies S1 to S3 View/request a protocol for this paper from Bio-protocol. Submitted 3 March 2023; accepted 5 June 2023 10.1126/science.adh1411 Xing et al., Science 381, 70–75 (2023) 7 July 2023 6 of 6
10.1126_science.adh2888
incarcerations that occurred in the country (16). Although homicide and incarceration trends are imperfect because they do not discriminate between offenses that occurred specifically in the context of organized crime, they can be used to estimate cartels’ violence capacity and the state’s incapacitation against them. In this work, we build on this intuition and exploit data on murders, missing persons, and incar- cerations in Mexico between 2012 and 2022 to derive cartel size. We propose a mathematical system to represent their behavior over 10 years and seek to shed light on the mechanisms with- in the so-called black box of the cartels. This work has two main goals. First, we aim to obtain plausible estimates of the cartels’ population, including their number of mem- bers and recruitment capacity. Second, we seek to simulate different policy scenarios (i.e., in- creased state incapacitation and recruitment prevention) to disentangle the effects of vary- ing strategies to curb cartels’ power and, in turn, violence in the country. Our conceptual framework is built on the evidence that, de- spite the high number of murders and incar- cerations in the past 10 years, cartels have maintained and even increased their power, control, and resources, introducing even more violence in the country. To construct our model, we gauge data on 150 cartels active in Mexico in 2020, including information on their alliances and rivalries and data corresponding to homi- cides, missing persons, and incarcerations. Methods We ask two research questions (RQs). RQ1: What is the size of Mexico’s cartel population, and what is their capacity to recruit members? RQ2: To control cartel violence, is a preventive policy strategy (focused on reducing cartel re- cruitment efforts) more effective than a reactive policy strategy (focused on increasing police efforts to incarcerate cartel members)? We consider four mechanisms that explain why cartel size varies: recruitment, incapac- itation, saturation, and conflict (Fig. 1). We model the conflict between cartels with a RES EARCH CARTELS Reducing cartel recruitment is the only way to lower violence in Mexico Rafael Prieto-Curiel1*, Gian Maria Campedelli2†, Alejandro Hope3‡ Mexican cartels lose many members as a result of conflict with other cartels and incarcerations. Yet, despite their losses, cartels manage to increase violence for years. We address this puzzle by leveraging data on homicides, missing persons, and incarcerations in Mexico for the past decade along with information on cartel interactions. We model recruitment, state incapacitation, conflict, and saturation as sources of cartel size variation. Results show that by 2022, cartels counted 160,000 to 185,000 units, becoming one of the country’s top employers. Recruiting between 350 and 370 people per week is essential to avoid their collapse because of aggregate losses. Furthermore, we show that increasing incapacitation would increase both homicides and cartel members. Conversely, reducing recruitment could substantially curtail violence and lower cartel size. out tasks (both strictly criminal and not) for cartels (11). Incapacitation measures the abil- ity of the state to counter cartels through in- carceration (12). Considering all incarcerations allows us to avoid the bias of only focusing on incarcerations for homicides, which are only a fraction of the offenses committed by cartel members. Conflict describes the extent to which cartels clash and fight with each other (13, 14). Finally, saturation character- izes internal instability and dropouts, which lead to organizational fragmentation (4, 15). Despite Mexican cartels’ economic, social, and political importance, we lack essential in- formation to better understand how they func- tion. In fact, we primarily lack estimates of the size of these criminal entities. We also lack systematic estimates of cartel-related killings and kidnappings and figures related to recruit- ment trends, which makes it extremely difficult to deepen our knowledge about their presence, resources, and goals. The secretive nature of cartels’ actions, as well as the insufficient amount of information accessible to map them, makes them conceptually similar to black boxes, from which we can only extrapolate imperfect proxies of activity using, for instance, the daily num- ber of homicides or the number of drug-related L atin America is home to only 8% of the world’s population, but roughly one in three intentional homicides worldwide occur in the region (1). Mexico accounts for a relevant share of such homicides, primarily because of the long-standing pres- ence of cartels across many areas of the coun- try. In 2021, Mexico reported 34,000 victims of intentional homicide—nearly 27 victims per 100,000 inhabitants—and was ranked among the least peaceful countries in Latin America (2). Between 2007 and 2021, the number of homicides in the country increased by more than 300% (3), with institutional sources quan- tifying that between 2006 and 2018, about 125,000 to 150,000 homicides were related to organized crime in Mexico (4). The effects of cartels on Mexico’s society are far-reaching. These entail their presence across a wide array of illegal activities beyond drug trafficking (5, 6), the deterioration of human rights (7), and the weakening of institutional stability through extensive acts of violence (8, 9). Furthermore, some cartels have acquired a transnational dimension, expanding their business to the United States and beyond (10). In this context, although cartels lose dozens of members daily as a result of killings and state incapacitation through incarcerations, the violence over the years has not decreased. We tackle this puzzle by studying cartels’ evo- lution, deriving their sizes, and considering four fundamental sources of size variation: recruitment, incapacitation, conflict, and sat- uration. These sources capture the different exogenous and endogenous dynamics explain- ing why and to what extent cartels grow or shrink. Recruitment refers to the process of attracting a new workforce that stably carries 1Complexity Science Hub, Vienna, Austria. 2Department of Sociology and Social Research, University of Trento, Trento, Italy. 3Independent Security Analyst. *Corresponding author. Email: prieto-curiel@csh.ac.at †Present address: Mobile and Social Computing Lab, Fondazione Bruno Kessler, Trento, Italy. ‡Deceased. Fig. 1. Model diagram representing the four reasons why a cartel changes in size. Most cartel-related activities remain undercover, but we observe some of their by-products in casualties and incapacitations. recruitment conflict 1500 weekly rate 1000 500 incapacitations missing persons recruitment + incapacitation − saturation − conflict 2012 2017 casualties 2022 − n w o n k n u n w o n k Prieto-Curiel et al., Science 381, 1312–1316 (2023) 22 September 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E weighted network, where a node represents each cartel, and an edge represents a conflict in some state in Mexico. Similarly, we con- struct a weighted network of alliances between cartels across different states. The model is a system of coupled differential equations, one for each cartel. Although we cannot observe most aspects of cartels (such as their recruitment and internal conflicts), we use the observed number of casualties and incarcerations to estimate the model parameters and infer the size of each cartel. We then use those esti- mates to forecast different scenarios for the next 5 years in Mexico. See the supplemen- tary materials, sections A to F, for details on methodology. Results RQ1: Estimating cartels’ populations Most cartel-related activities are organized as dark networks to maintain their operations and activities covered (17, 18). However, their human losses caused by homicidal violence and the state’s action through incapacitation provide insights into the overall amounts of such activities. We leverage the trends in homi- cides, missing persons, and incarcerations over the past decade to motivate our investiga- tion of cartels’ sizes in Mexico (supplementary materials, section A). Not all losses are directly related to the conflict between cartels (e.g., domestic violence), and some are a by-product of their dispute (e.g., deaths suffered by family members or bystanders). To study the size and evolution of the cartel population, we exclu- sively model homicides between cartel mem- bers (i.e., homicides in which the victim and the perpetrator are both cartel members). The starting point is that cartels have not seen their power diminished because violence has not re- duced either. In Mexico, 686 people were killed each week of 2021, with an additional 137 people reported as missing and yet to be found, and more than 2500 people were imprisoned each week (3, 19, 20). We use the number of cartel losses to infer otherwise unknown properties, including their size and recruitment rate. Data compiled from open sources by the Programa de Política de Drogas (PPD) (21) enable us to detect the ex- istence of k = 150 active cartels in Mexico in 2020. Building on such data, we operationally define cartels as those criminal organizations that are found to be active in Mexico, regard- less of their size and activity (supplementary materials, section B). Cartels have different interactions: They can be allies, they can have no interactions (particularly from distant lo- cations), or they can fight for territory or re- sources, creating substantial losses among both groups. To represent these interdepen- dencies, we construct two separate weighted networks—the allies A and rivalries R—to recreate conflicting and cooperating cartels, Fig. 2. Rivalries and alliances were observed between 150 active cartels in Mexico in 2020. The size of the node represents the estimated cartel size. If cartels have at least one state rivalry, nodes are connected (left). The width of the edge corresponds to the number of states in which cartels fight. Nodes are connected if they are identified as allies (right). NF Mich., Nueva Familia Michoacana; UTepito, Unión Tepito; Z, Los Zetas; SRdL, Santa Rosa de Lima. with weights corresponding to the number of states in which two cartels interact (Fig. 2). Major cartels, such as Cártel Jalisco Nueva Generación (CJNG), the Sinaloa Cartel, and Nueva Familia Michoacana, are present al- most at a national level and have alliances with many satellite organizations forming three main clusters. These clusters fight against each other, creating most of the violence between cartels (16). Smaller organizations are local to one city and have few interactions (cooperation or con- flict) with other cartels. X The number of members of cartel i at time t, expressed as Ci(t), increases instantly accord- ing to rCi, where r is the fixed recruitment rate. Because of state forces, the size of the cartel decreases by hCi= jCj for some h > 0 that represents the incapacitation rate. Because of internal instability, dropouts, and dimin- ishing returns, large groups decrease their size instantly by wC 2 i for some small value of w > 0, known as the saturation rate (22, 23). The impact of conflict between two cartels, i and j, is modeled according to the number of homicide offenders between rival groups, which is assumed to be proportional to car- tel size, so cartel i suffers instant casualties according to qCiCj, where q ≥ 0 is the deathly rate of conflict related to homicide offend- ers within cartels. Combining recruitment, incarceration, conflict, and saturation, we obtain (cid:2) C (cid:3) h i ¼ rCi |{z} recruitment Ci C|{z} incapacitation Xk (cid:3) q CiCjSij j≠1 |{z} conflict ð1Þ (cid:3) wC 2 i |{z} saturation (cid:2) where C i indicates the rate of change in cartel size i, and Sij ≥ 0 captures the interaction be- tween cartels. We obtain a system of k = 150 coupled differential equations—one for each cartel (supplementary materials, section C). The number of weekly casualties produced by all cartels is given by d tð Þ ¼ qC⊤SC, where C = (C1,C2,…,Ck). Cartels recruit rC individuals, where C ¼ Ci, and i tð Þ ¼ h are inca- Xk X i¼1 Ci C pacitated. In line with previous works on other types of organizations, we assume that the ini- tial cartel size is a heavy-tailed distribution (sup- plementary materials, section D) (24–26). We use the observed weekly number of casualties and incapacitations to estimate the time-varying number of cartel members Ci(t). Not all observed deaths, missing persons, and incapacitations in the country are suffered by cartel members, and most incapacitations are not linked to the incarcerations of cartel members. In our analysis, we estimate casual- ties as the sum of missing persons with mur- ders and consider that a fraction f = 10% of the observed weekly deaths and a fraction g = 5% of the incapacitations are cartel members (supplementary materials, section D). In total, 50,000 casualties and 55,000 incapacitations directly involve cartel members. On the basis of these figures, we estimate that in 2012, there were 115,000 cartel members and that by 2022, the number increased to 175,000. Thus, despite efforts from the state to hinder their power, car- tels have increased their size by 60,000 members in a decade. Incarcerating nearly 6000 cartel members each year has not prevented them from growing into larger organizations. Given the current conditions, we quantify 120 weekly cartel-related deaths, with an increase of 77% between 2012 and 2022. To ensure that our results are not driven by wrong assumptions Prieto-Curiel et al., Science 381, 1312–1316 (2023) 22 September 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E about the number of homicides between cartel members and incarcerations of cartel affili- ates, we conduct sensitivity tests considering the scenarios between 40,000 and 60,000 car- tel casualties and 45,000 and 65,000 incapa- citations. By considering the variation of these two parameters, we obtain that the total pop- ulation of cartel members in 2022 lies between 160,000 and 185,000 units. At the same time, additional sensitivity tests were used to try to quantify the effect of potential missing data at the network level concerning alliances and rivalries. Adding 10% more cartels would, on average, lead to 3.2% more members than the estimated 175,000. Furthermore, we also pro- vide evidence that adding 10% more alliances or rivalries would at most affect the overall dimension of violence by 5% (supplementary materials, section E). Even under a conserva- tive scenario, Mexican cartels have lost around 200 members per week for years (Fig. 3A). Spe- cifically, we estimate that in a decade, 285,000 people acted as cartel units and that—in total— 37% of them are either deceased (17%) or in- carcerated (20%). Despite competition with other cartels and state forces’ incapacitation, cartels have prevailed for decades. Between January and December of 2021, cartels recruited 19,300 individuals, losing 6500 members as a result of conflict with other cartels and 5700 members as a result of incapacitation, which resulted in a net gain of roughly 7000 members during that year (supplementary materials, section D). A similar estimate is observed for each year be- tween 2012 and 2022. Unless all cartels com- bined recruit between 350 and 370 people per week, they would have collapsed as a re- sult of conflict, incapacitation, and saturation combined (Fig. 3A). Given the estimated overall population, all cartels combined are the fifth largest employer in Mexico (27) (Fig. 3B). The 10 largest cartels in Mexico have more than 50% of the active affiliates in the country, but the conflict be- tween them only produces 15% of the fatalities (Fig. 3C). Most cartels are small local organ- izations playing a critical role in creating vi- olence in the country, often becoming targets of more powerful organizations. Previous re- search has suggested that large cartels fre- quently adopt fragmented cells of other weaker and less experienced structures (16). Small car- tels play a crucial role because they are more likely to become targets of powerful illicit organizations rather than fighting organ- izations of similar sizes. We estimate that more than half of the country’s casualties result from the fight between the smallest 140 and the largest 10 cartels (supplemen- tary materials, section B). RQ2: Comparing policy scenarios On the basis of the size of cartels in 2022 and the trends observed in the past decade, we predict that the weekly number of casualties related to organized crime will keep increasing in the coming years. We estimate that if current trends continue, cartels will keep increasing their power, and we could observe 40% more casualties and 26% more cartel members by 2027. We test the effectiveness of two main pol- Fig. 3. Current size of cartels and career paths for recruited members. (A) Between 2012 and 2022, we estimate that 285,000 people took part as cartel members, but only 60% were still active by 2022. The cartel career is brief and risky. Roughly 17% of them are dead, and 20% are incapacitated. (B) Number of employees from the top 10 companies in Mexico and the combined size of cartels (27). We estimate that cartels had between 160,000 and 185,000 members combined. (C) Of the 175,000 active cartel members, roughly 17.9% are part of CJNG, 8.9% are part of Cartel de Sinaloa, and 6.2% are from Nueva Familia Michoacana—the top three cartels in size. icy scenarios designed to reduce future violence in the country: first, a preventive strategy aimed at reducing cartel recruitment, and second, a reactive strategy aimed at increasing incapac- itation. On the one side, doubling incapacitation, with all of the associated costs and challenges in increasing security resources (including po- lice personnel, army, prisons, etc.), will still result in an increase of 8% in the number of casualties and an increase of 6% in the num- ber of cartel members. Even doubling incar- cerations will translate to a rise in violence (Fig. 4). Cartels have a critical equilibrium where their recruitment compensates for their losses, maintaining a stable size. Yet, if the recruitment rate of a cartel is 10% above its equilibrium, the incapacitation rate must increase by more than 21% to dismantle it (supplementary ma- terials, section F). Conversely, decreasing the cartel’s ability to recruit by half will reduce the weekly casualties by 2027 by 25% and cartel size by 11%. Math- ematically, a preventive strategy is far more successful than a traditional reactive strategy. However, the cartel population is so large that, even in the hypothetical scenario where recruit- ment drops to zero, it would take 3 years to return to the—already high—levels of violence observed in 2012. This further calls for rapid and timely large-scale initiatives to reduce re- cruitment in the country. We also assess the effects of two additional ancillary policy scenarios. The first one is de- signed to alter the type of conflict between car- tels (e.g., pushing for a narcopeace), and the second one is targeted at modifying cartels’ saturation levels (i.e., making cartels more prone to fragmentation). Neither of the two strat- egies outperforms the positive effects that a reduction in recruitment could produce (sup- plementary materials, section E). Decreasing the conflict by 20% reduces the number of ca- sualties by 8.7%, whereas increasing satura- tion by 20% lowers the number of homicides between cartel members by 5.4% (supplemen- tary materials, section E). In light of the cur- rent estimated circumstances, the growth of cartels’ size is impeded mainly by the conflict existing among organizations rather than the ability of the state to reduce the levels of vi- olence in Mexico successfully. Discussion For the past 15 years, Mexico has suffered from staggering levels of violence. Most of the vi- olence has been perpetrated by cartels fighting against each other (4). Despite the relevance of cartels, we lack basic information on their size and the impact of different policies that seek to curb their power. To the best of our knowl- edge, this work represents the first scholarly attempt to mathematically quantify the size of the cartel population in Mexico and to compare policy scenarios intended to decrease violence Prieto-Curiel et al., Science 381, 1312–1316 (2023) 22 September 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Forecast of the number of casualties and cartel size according to four different strategies. Weekly cartel-related deaths (top) and cartel size (bottom) if trends continue, if incapacitation doubles, if recruitment is reduced by half, and if recruitment is reduced to zero. Estimates for 2027 are obtained by keeping the 2022 estimates and adjusting the corresponding values of incapacitation or recruitment. in the country. Overall, our work advances the growing literature on mathematical and sta- tistical simulations for studying complex crim- inal phenomena (28–30). Our simulations yield some key findings. We estimated that the cartel population counted 160,000 to 185,000 units by 2022 and that, over the 2012 to 2022 period, 285,000 people acted as cartel members. Given these figures, we showed that in 2022, cartels needed to recruit between 350 and 370 units per week to avoid collapse as a result of joint effects of conflict, incapacita- tion, and saturation. Furthermore, we assessed the effectiveness of two main scenarios to curb cartels’ violence: preventive (intended to pre- vent recruitment) and reactive (designed to in- crease incapacitation through incarcerations). If current levels of incapacitation are doubled, some violence will be contained, but we would still expect an increase in the weekly casualties. Conversely, reducing recruitment by half leads to a decrease in homicides of 25%. We also tested the effect of two ancillary scenarios—reducing the conflict by pushing for cartel agreement and fragmentation, intended to decrease car- tels’ power through internal fights (supplemen- tary materials, section E). Results showed that the preventive strategy remained substantially more effective in reducing violence in the coun- try. Tackling recruitment will have a triple effect in the future: First, it will lower the number of cartel members, reducing the violence that it can create by having fewer killers. Second, it will lower the number of targets, so fewer people are vulnerable to suffering more violence. And third, it will reduce the cartel’s capacity for fu- ture recruitment. Although offering policy recommendations is beyond the scope of this work, our results can prompt policy-related reflections. Many initiatives to counter organized crime aim to increase incapacitation through incarceration. In this work, we demonstrate how increasing incapacitation substantially may not necessarily reduce violence. Contrarily, we offer an alterna- tive scenario centered around reducing recruit- ment and suggest how it may have longer-lasting beneficial effects. More than 1.7 million people in Latin America are incarcerated, and adding more people to saturated jails will not solve the insecurity problem (31). Despite the contributions of this investigation, there were some limitations. First and foremost, although the lack of data on the size of cartels represents the inherent motivation of this work, it also represents a structural limitation because our estimates cannot be meaningfully validated with real-world information. We took all possi- ble precautions to obtain statistically consistent estimates through extensive sensitivity analyses, but this does not eliminate the core validation issue. Additionally, a thorough reflection on other sources of limitation and assumptions is provided in the supplementary materials, sec- tion I. These entail (i) temporal variability in rivalries and alliances, (ii) alternative sources of cartels’ size variability, and (iii) the lack of a finite population. Results highlight the need to devote more attention to recruitment. Reducing recruit- ment requires structural efforts at the state and local levels. This especially applies to areas with high cartel support, where offering edu- cational and professional opportunities that outweigh the short-term benefits offered by cartels represents a critical goal for the future of the country (32–35). Future work on this topic should focus on enriching our model of cartel size variation with additional sources, such as cartel fragmentation, and should also consider the possibility of studying recruit- ment dynamics using data on finite populations to obtain mathematical models that consider individual risk factors (such as age and sex) in the computation of violence and recruit- ment trends. REFERENCES AND NOTES 1. UNODC, UN Office on Drugs and Crime’s International 2. 3. Homicide Statistics Database (2010). Institute for Economics & Peace, “Global Peace Index 2022: Measuring Peace in a Complex World” (2022); https://www. visionofhumanity.org/wp-content/uploads/2022/06/ GPI-2022-web.pdf. Instituto Nacional de Estadística y Geografía INEGI, Defunciones por homicidio 1990-2021 (deaths by homicide, 1990-2021) (2022); https://www.inegi.org.mx/temas/mortalidad/. 4. Congressional Research Service, “Mexico: Organized Crime and Drug Trafficking Organizations” (CRS report R41576, 2020); https://crsreports.congress.gov/product/pdf/R/R41576/45. 5. N. P. Jones, J. P. Sullivan, J. Strateg. Secur. 12, 1–24 (2019). 6. F. Carbajal Glass, in Illegal Mining: Organized Crime, Corruption, and Ecocide in a Resource-Scarce World, Y. Zabyelina, D. van Uhm, Eds. (Palgrave Macmillan Cham, 2020), pp. 147–183. 7. A. Anaya-Muñoz, B. Frey, Mexico’s Human Rights Crisis (Univ. of Pennsylvania Press, 2018). 8. A. Trelles, M. Carreras, J. Polit. Lat. Am. 4, 89–123 (2012). 9. B. J. Phillips, Perspect. Terror. 12, 46–63 (2018). 10. Europol, US Drug Enforcement Administration (DEA), “Complexities and conveniences in the international drug trade: The involvement of Mexican criminal actors in the EU drug market” (2022); https://www.europol.europa.eu/cms/ sites/default/files/documents/Europol_DEA_Joint_Report.pdf. 11. F. Calderoni et al., Campbell Syst. Rev. 18, e1218 (2022). 12. J. Cohen, Crime Justice 5, 1–84 (1983). 13. G. Trejo, S. Ley, Comp. Polit. Stud. 51, 900–937 (2018). 14. L. H. Atuesta, A. F. Ponce, Glob. Crime 18, 375–402 (2017). 15. L. H. Atuesta, Y. S. Pérez-Dávila, Trends Organ. Crime 21, 235–261 (2018). 16. N. P. Jones, I. Chindea, D. Weisz-Argomedo, J. P. Sullivan, “Mexico’s 2021 dark network alliance structure: An exploratory social network analysis of Lantia Consultores’ illicit network alliance and subgroup data” (Rice University’s Baker Institute for Public Policy, 2022); https://doi.org/10.25613/KMGB-NC83. 17. R. M. Bakker, J. Raab, H. B. Milward, J. Policy Anal. Manage. 31, 33–62 (2012). 18. J. Raab, H. B. Milward, J. Public Adm. Res. Theory 13, 413–439 (2003). 19. Comisión Nacional de Búsqueda de Personas, El Registro Nacional de Personas Desaparecidas y No Localizadas (RNPDNO) (Gobierno de México, 2022); https:// versionpublicarnpdno.segob.gob.mx/. 20. INEGI, Censo Nacional de Sistema Penitenciario Federal 2021 (2020); https://www.inegi.org.mx/programas/cnspef/2021/ #Datos_abiertos. 21. Política de Drogas México, Plataforma de Proyección de Datos Abiertos (2022); https://ppdata.politicadedrogas.org/#ppd.gc. 22. R. H. Coase, Economica 4, 386–405 (1937). 23. J. P. Caulkins, G. Feichtinger, D. Grass, G. Tragler, Oper. Res. Lett. 37, 387–391 (2009). 24. A. Clauset, C. R. Shalizi, M. E. Newman, SIAM Rev. 51, 661–703 (2009). 25. N. F. Johnson et al., Universal patterns underlying ongoing wars and terrorism. arXiv:physics/0605035 [physics.soc-ph] (2006). 26. D. D. J. Restrepo, M. Spagat, S. van Weezel, M. Zheng, N. F. Johnson, J. Comput. Sci. 46, 101088 (2020). 27. Expansión, “Las 500 empresas más importantes de México” (2022); https://expansion.mx/las-500-empresas-mas-importantes-mexico. 28. G. M. Campedelli, Machine Learning for Criminology and Crime Research: At the Crossroads (Routledge, 2022). 29. F. Calderoni, G. M. Campedelli, A. Szekely, M. Paolucci, G. Andrighetto, J. Quant. Criminol. 38, 197–237 (2022). 30. D. Weisburd, M. Wolfowicz, B. Hasisi, M. Paolucci, G. Andrighetto, Criminol. Public Policy 21, 461–485 (2022). 31. H. Fair, R. Walmsley, Institute for Crime & Justice Policy Research, World Prison Brief (2022); https://www.prisonstudies.org/. 32. T. E. Murphy, M. A. Rossi, J. Dev. Econ. 143, 102433 (2020). 33. N. P. Jones, J. Strateg. Secur. 11, 19–42 (2018). 34. S. D. Levitt, S. A. Venkatesh, Q. J. Econ. 115, 755–789 (2000). 35. B. J. Phillips, V. Ríos, Lat. Am. Polit. Soc. 62, 1–24 (2020). 36. R. Prieto-Curiel, G. M. Campedelli, Mexican cartels form a network of alliances and rivalries, dataset, Dryad (2023); https://doi.org/10.5061/dryad.zw3r228d7. AC KNOWLED GME NTS We are grateful for the insightful comments from L. Sánchez, J. Mohar, and C. A. Pérez Ricart. After submitting this manuscript to Science, author Alejandro Hope passed away. We acknowledge his tireless efforts to make Mexico a safer and more peaceful country, and we hope that this work will honor his memory. Funding: The research was funded by the Austrian Federal Ministry for Climate Action, Environment, Energy, Mobility, Innovation and Technology Prieto-Curiel et al., Science 381, 1312–1316 (2023) 22 September 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E (2021-0.664.668) and by the Austrian Federal Ministry of the Interior (2022-0.392.231). Author contributions: R.P.-C. designed the study. R.P.-C. and G.M.C. analyzed the results. All authors wrote the manuscript, but A.H. passed away while it was under peer review. R.P.-C. and G.M.C. completed the revisions. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data are available in the manuscript or the supplementary materials. A public repository with the processed data and code to reproduce and extend the results is available on Dryad (36). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal-article-reuse Figs. S1 to S8 Tables S1 and S2 References (37–66) MDAR Reproducibility Checklist Spanish Translation of Author Accepted Manuscript SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh2888 Supplementary Text Submitted 22 February 2023; accepted 2 August 2023 10.1126/science.adh2888 Prieto-Curiel et al., Science 381, 1312–1316 (2023) 22 September 2023 5 of 5
10.1126_science.adh3428
RES EARCH R E S E A R C H A R T I C L E ◥ COMPARATIVE COGNITION Songbird species that display more-complex vocal learning are better problem-solvers and have larger brains Jean-Nicolas Audet1,2*, Mélanie Couture1,3, Erich D. Jarvis1,2,3,4* Complex vocal learning, a critical component of human spoken language, has been assumed to be associated with more-advanced cognitive abilities. Tests of this hypothesis between individuals within a species have been inconclusive and have not been done across species. In this work, we measured an array of cognitive skills—namely, problem-solving, associative and reversal learning, and self-control—across 214 individuals of 23 bird species, including 19 wild-caught songbird species, two domesticated songbird species, and two wild-caught vocal nonlearning species. We found that the greater the vocal learning abilities of a species, the better their problem-solving skills and the relatively larger their brains. These conclusions held when controlling for noncognitive variables and phylogeny. Our results support a hypothesis of shared genetic and cognitive mechanisms between vocal learning, problem-solving, and bigger brains in songbirds. complex vocal learning behavior overlap with those long thought to exhibit more-intelligent cognitive capacities [e.g., humans, cetaceans, elephants, corvid songbirds, and parrots (3, 4)], although this has not been quantitatively tested across species. The few studies that have tested for rela- tionships between vocal learning complexity and cognitive abilities have all been within species (5). Some studies found positive rela- tionships (6, 7), whereas others did not (8–10), and some found negative relationships (7, 11), which led Searcy and Nowicki (5) to conclude that there is not good evidence of such relation- ships. The challenge for these within-species studies is that it is difficult to tease out whether individual differences are due to genetic fac- tors or cultural or life-experience factors, such as food deprivation (12) or stress (13). Although vocal learning is often thought of as a dichotomous trait, considerable variation in phenotypes has been documented within lineages, with songbirds being the most speciose S poken language and problem-solving are often considered to be components of intelligence in humans. An essential and specialized component of spoken language is vocal production learning, or the ability to imitate sounds (1, 2). Ad- vanced vocal learning has been found in only a handful of taxa, including five mammalian (humans, elephants, cetaceans, pinnipeds, and bats) and three avian (songbirds, parrots, and hummingbirds) clades (1). Interestingly, the vocal learning taxa that display the most- 1The Rockefeller University Field Research Center, Millbrook, NY, USA. 2Laboratory of Neurogenetics of Language, The Rockefeller University, New York, NY, USA. 3The Vertebrate Genome Laboratory, The Rockefeller University, New York, NY, USA. 4Howard Hughes Medical Institute, The Rockefeller University, New York, NY, USA. *Corresponding author. Email: jaudet@rockefeller.edu (J.-N.A.); ejarvis@rockefeller.edu (E.D.J.) Fig. 1. The 23 study species of birds and their vocal learning characteristics. (A) Phylogenetic tree of the species that were studied based on (42). Inside circle colors indicate the type of vocal learning, perimeter circle colors indicate domesticated versus wild species, and circle size is proportional to vocalization repertoire size. (B) Comparisons of song and call repertoire sizes in open- versus closed-ended vocal learning songbirds (n = 21 species). (C) Comparison of mimic and nonmimic species. Black * and ** indicate significant PANOVA at <0.05 and <0.01, respectively. Green * and ** indicate significant PAOV.PHYLO (ANOVA accounting for phylogenetic relationships). ns, not significant. Full statistical values are provided in table S2. Bars are means ± SEM. Audet et al., Science 381, 1170–1175 (2023) 15 September 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Table 1. Summary of MCMCglmm phylogenetic models. Full, separate models were implemented for vocal learning type [closed-ended (reference category), open-ended, and mimicry]; repertoire size of songs, calls, or both; and vocal learning complexity. All potential covariates were tested first (see tables S4 to S7), and the models with only the significant variables, when applicable (with species, phylogeny, and capture site as random effects), were rerun to obtain unbiased effects of vocal learning. Significant effects are highlighted in bold. All measured behaviors are expressed in logged trials; higher numbers represent lower performance (n = 23 species, 214 individuals). Post.mean, mean of the posterior distribution; CI, confidence interval; pMCMC, MCMCglmm P value. Cognitive trait Vocal learning feature Post.mean Lower 95% CI Upper 95% CI pMCMC Learning Problem-solving Reversal learning 0.0020 Open-endedness ......................................................................................................................................................................................................................................................................... 0.0378 Mimicry ......................................................................................................................................................................................................................................................................... 0.0217 Song repertoire ......................................................................................................................................................................................................................................................................... 0.0216 Call repertoire ......................................................................................................................................................................................................................................................................... 0.0018 Total vocal repertoire ......................................................................................................................................................................................................................................................................... 0.0012 Vocal learning complexity ............................................................................................................................................................................................................................................................................................................................................ 0.0889 Open-endedness ......................................................................................................................................................................................................................................................................... 0.3065 Mimicry ......................................................................................................................................................................................................................................................................... 0.6366 Song repertoire ......................................................................................................................................................................................................................................................................... 0.4620 Call repertoire ......................................................................................................................................................................................................................................................................... 0.8780 Total vocal repertoire ......................................................................................................................................................................................................................................................................... 0.3768 Vocal learning complexity ............................................................................................................................................................................................................................................................................................................................................ 0.0269 Open-endedness ......................................................................................................................................................................................................................................................................... 0.1634 Mimicry ......................................................................................................................................................................................................................................................................... 0.4115 Song repertoire ......................................................................................................................................................................................................................................................................... 0.7206 Call repertoire ......................................................................................................................................................................................................................................................................... 0.9675 Total vocal repertoire ......................................................................................................................................................................................................................................................................... 0.1978 Vocal learning complexity ............................................................................................................................................................................................................................................................................................................................................ 0.7938 Open-endedness ......................................................................................................................................................................................................................................................................... 0.2026 Mimicry ......................................................................................................................................................................................................................................................................... 0.9554 Song repertoire ......................................................................................................................................................................................................................................................................... 0.8932 Call repertoire ......................................................................................................................................................................................................................................................................... 0.9742 Total vocal repertoire ......................................................................................................................................................................................................................................................................... 0.4057 Vocal learning complexity ............................................................................................................................................................................................................................................................................................................................................ 0.1571 Open-endedness ......................................................................................................................................................................................................................................................................... 0.0812 Mimicry ......................................................................................................................................................................................................................................................................... 0.0362 Song repertoire ......................................................................................................................................................................................................................................................................... 0.0132 Call repertoire ......................................................................................................................................................................................................................................................................... 0.0013 Total vocal repertoire ......................................................................................................................................................................................................................................................................... 0.0037 Vocal learning complexity ............................................................................................................................................................................................................................................................................................................................................ −1.151 −1.036 −0.174 −0.187 −0.257 −0.090 −0.391 −0.295 0.024 −0.038 −0.009 −0.016 −0.818 −0.607 0.066 −0.029 0.005 −0.037 −0.111 −0.725 0.003 −0.009 0.002 −0.023 0.976 1.971 1.373 1.702 2.297 0.678 −0.517 −0.056 −0.030 −0.032 −0.118 −0.046 0.060 0.297 0.126 0.068 0.106 0.021 −0.133 0.272 0.227 0.140 0.185 0.021 0.667 0.443 0.148 0.147 0.169 0.034 2.362 4.242 2.628 2.989 3.494 1.119 −1.806 −2.015 −0.323 −0.342 −0.395 −0.134 −0.843 −0.890 −0.078 −0.141 −0.124 −0.054 −1.497 −1.469 −0.092 −0.194 −0.173 −0.093 −0.910 −1.928 −0.144 −0.162 −0.160 −0.080 −0.390 −0.248 0.114 0.392 1.125 0.242 Relative brain size Self-control and best-characterized clade (1, 14–16). Features that are believed to reflect more-complex vocal learning in songbirds include (i) large vocal repertoires, (ii) lifelong open-ended vocal learn- ing ability, and (iii) mimicry of other species’ vocalizations (16–18). It is unknown whether these vocal learning variations are linked with other cognitive phenotypes. Cognition has been measured in the labo- ratory using a variety of behavioral tasks. Of these, problem-solving is assumed to be one of the most complex, requiring animals to fig- ure out the best action to overcome a challenge. Problem-solving may be relevant for allow- ing animals to cope with ecological disturbances (19). Other cognitive assays that are often con- ceptually linked with intelligence include asso- ciative learning, reversal learning, and self-control tasks (20). In this work, we quantitatively tested whether there is a linkage between vocal learning com- plexity and other cognitive traits. We tested 214 individuals from 23 species consisting of 21 wild-caught and two domesticated avian spe- cies; 21 species were vocal learning songbirds, and two were vocal nonlearning species (Fig. 1A). We generated a database of vocal learning char- acteristics of these species from an extensive literature of 80 publications (table S1), assessed species’ performance on a battery of behav- ioral tasks (fig. S1 and supplementary meth- ods), and found across-species relationships between vocal learning complexity, problem- solving, and relative brain size. Vocalization repertoire size is associated with open-endedness and mimicry Wild birds were caught at the Rockefeller Uni- versity Field Research Center in New York, USA, over a 3-year period, and domesticated species were bred at the same location (zebra finches) or purchased from a local breeder (canaries). Birds were opportunistically cap- tured in mist nets, and species were chosen if their local abundance allowed for a sufficient sample size (n ≥ 12) of males (the vocal learn- ing sex for most species); in addition, for eight species, we kept and tested one or two animals each to see whether they would follow a trend. All birds were habituated individually in cages for 3 days, where they could hear but not see other individuals. For all 23 species, literature exists on their vocal behavior (table S1). We used these data to establish a consistently defined profile for each species in a database that included six vocal learning characteristics: (i) presence of vocal learning, (ii) open-ended versus closed-ended vocal learning, (iii) capacity for vocal mimicry of other species, (iv) song repertoire size per individual, (v) call repertoire size per species, and (vi) total repertoire size of both songs and calls (supplementary methods and table S1). Audet et al., Science 381, 1170–1175 (2023) 15 September 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E ) k n a r ( g n i v l o s - m e b o r P l ) k n a r ( i g n n r a e l e v i t a i c o s s A ) k n a r ( g n n r a e l i l a s r e v e R ) k n a r ( l o r t n o c - f l e S 5 10 15 20 5 10 15 20 5 10 15 20 5 10 15 20 Vocal learning types Vocalization repertoire size A *** B ns I *** Non-learning Closed-ended Open-ended Mimic P = 0.0003 P = 0.1389 C ns D ns J P = 0.1858 P = 0.6506 E * F ns K P = 0.0162 P = 0.5029 G ns H ns L P = 0.8253 P = 0.7491 Closed- ended Open- ended Non- mimic Mimic Blue jay (n=13) Tufted titmouse (n=13) Black-capped chickadee (n=19) Cedar waxwing (n=2) White-breasted nuthatch (n=13) House wren (n=13) Gray catbird (n=15) European starling (n=16) American robin (n=12) Veery (n=1) Zebra finch (n=12) American goldfinch (n=13) Canary (n=13) Yellow warbler (n=1) American redstart (n=2) Red-winged blackbird (n=2) Brown-headed cowbird (n=14) Chipping sparrow (n=12) White-throated sparrow (n=12) Song sparrow (n=1) Northern cardinal (n=1) Eastern phoebe (n=13) Mourning dove (n=1) R = 0.814 P < 0.0001 ns R = -0.171 P = 0.4465 ns R = 0.059 P = 0.7951 ns R = -0.016 P = 0.9429 20 Vocalization repertoire size (rank) 15 10 5 Fig. 2. Relationships between cognitive traits and vocal learning features. (A to H) Four cognitive measures (y axes) compared between species grouped by vocal learning features (x axes: closed-ended, which includes vocal nonlearning species; open-ended; and nonmimic or mimic). (I to L) Correlation analyses between the four cognitive behavioral measures (y axes) and vocalization repertoire size (x axis). Values are ranked means of species performance from all individuals (sample sizes are shown in the legend). P values in (A) to (H) are from Wilcoxon tests; R and P values in (I) to (L) are from Spearman correlations. Regression lines are for illustration purposes to show the significance and direction of relationships. Images to the left are examples of birds in the different tasks, which are snapshots from videos taken by the authors. ***P < 0.001; *P < 0.05; ns, not significant. We found that, consistent with Robinson et al. (18), open-ended vocal learning songbirds had significantly larger song repertoires than closed- ended vocal learners (Fig. 1B and table S2A). Using the complete vocalization repertoires (song and calls) yielded an even greater dif- ference between open- and closed-ended vocal learning species (Fig. 1B and table S2A). Similarly, songbird species that are capable of vocal mimicry had larger reper- toires than nonmimics (Fig. 1C and table S2A). All of these differences increased in significance when including the two vocal nonlearners (the mourning dove and suboscine eastern phoebe; table S2B), which were at the lower end of the distribution for song and call re- pertoire sizes (table S1). Nearly all differ- ences were still significant when accounting for phylogenetic relationships in the analy- sis of variance (ANOVA) model, except for repertoire size between mimic and nonmimic songbirds, which still approached significance [phylogenetic ANOVA P value (PAOV.PHYLO) ~ 0.07; table S2B]. These findings demonstrate that relationships exist among different vocal learning phenotypes. Species with more-advanced vocal learning abilities are better problem-solvers We presented all 214 individuals of the 23 bird species with seven cognitive tasks after over- night food deprivation, the duration of which was adjusted to the night lengths and body weights of each individual. The tasks occurred over 6 days in the same sequence, with 5-min intervals between trials or tasks (supplementary Audet et al., Science 381, 1170–1175 (2023) 15 September 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Species that display higher vocal learning complexity are better problem-solvers. (A) PCA on open-ended capacity, mimicry capacity, and log of vocalization repertoire (songs and calls). PC1 explains 69.9% of the variance and was used as our index of vocal learning complexity. (B) The higher the vocal learning complexity, the better the problem-solving performance among species. (C to E) Species’ associative learning (C), reversal learning (D), and self-control (E) performances are not associated with vocal learning complexity. Values are ranked means of species performance from all individuals (sample sizes are shown in the legend). R and P values are from Spearman correlations. ***P < 0.001; ns, not significant. methods). The first four were different obstacle- removal tasks of increasing complexity, during which the birds had to figure out how to ac- cess a food reward (seeds or worms) by either removing, piercing, or pulling parts of the apparatuses (fig. S1, F to I, and movies S1 to S4). We used the average number of trials required to solve the four problems as our problem-solving measure. Self-control was eval- uated using a typical detour-reaching task in which birds had to access a food reward without trying to obtain the food through a transparent barrier (fig. S1J and movies S5 to S6). Associative learning was measured on a standard two-color discrimination task in which birds had to associate a color with a food reward (fig. S1K and movie S7). The rewarded color was switched the next day to assess re- versal learning. Looking at the categorical vocal learning variables, we found that species classified as open-ended vocal learners were significantly better problem-solvers than the other species (Fig. 2A). Species capable of vocal mimicry were at the upper problem-solving performance range, although the difference was not signif- icant (Fig. 2B). There were no significant dif- ferences for open-ended or mimicking vocal learning species and their associative learning abilities (Fig. 2, C and D), reversal learning (Fig. 2, E and F), or self-control (Fig. 2, G and H), except a largely overlapping, but signif- icantly better, reversal learning performance in open-ended vocal learners (Fig. 2E). When splitting the vocal learning phenotypes into a more-sensitive multigroup ANOVA analysis, only problem-solving was significantly better for both open-ended vocal learners and mim- ics (fig. S2, A to D). Looking at the continuous vocal learning variables, we found a strong and significant relationship, with species having the largest repertoires (songs and calls) being the best problem-solvers (Fig. 2I). We found no signif- icant correlations or even qualitative signs of a correlation between vocalization repertoire size and associative learning, reversal learn- ing, or self-control (Fig. 2, J to L). Considering only songs or calls yielded similar significant relationships with problem-solving, although the associations were weaker (table S3). Exclud- ing vocal nonlearning and domesticated species yielded identical conclusions (fig. S2, F to I). When excluding the eight species with small sample sizes, the correlation between vocal- ization repertoire size and problem-solving re- mained (Spearman’s correlation R = 0.779; P = 0.0006). Together, these results indicate a posi- tive association between vocal learning features and problem-solving among cognitive traits. Vocal learning complexity better predicts problem-solving performance To obtain an estimate of general vocal learn- ing complexity, we performed a principal com- ponents analysis (PCA) with the three vocal learning features (open-endedness, mimicry, and total vocalization repertoire size) and ex- tracted the first principal component (PC1), which explained the majority (69.9%) of the variance in all three measures (Fig. 3A). We found a positive and significant relationship Audet et al., Science 381, 1170–1175 (2023) 15 September 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Relationships between relative brain size and vocal learning features. (A) Open-ended vocal learning species have significantly larger relative brain sizes than the other species. (B) The difference in brain size between mimic and nonmimic species is not significant. (C) Vocalization repertoire is significantly associated with relative brain size. (D) Species’ vocal learning complexity is also significantly associated with relative brain size. Relative brain sizes are the residuals of brain volumes with body size obtained from (43). P values in (A) and (B) are from Wilcoxon tests; R and P values in (C) and (D) are from Spearman corre- lations, with vocal learning complexity from PC1 of Fig. 3A. ***P < 0.001; *P < 0.05; ns, not significant. between the vocal learning complexity PC1 and problem-solving across species, which was stronger than with any of the individual vocal learning features (Fig. 3B versus Fig. 2, A to B and I). Vocal learning complexity was unrelated to the other cognitive traits of asso- ciative learning, reversal learning, or self-control (Figs. 3, C to E). Again, excluding vocal non- learning and domesticated species did not change the outcome of the results (fig. S3). When excluding the eight species with small sample sizes, the correlation between vocal learning complexity and problem-solving re- mained (R = 0.846; P = 0.0001). These findings strengthen the conclusion of an association between vocal learning abilities and problem- solving across species and provide a means to summarize overall vocal learning complexity in one measure. More-advanced vocal learners have bigger brains We sought a biological variable that could ex- plain our findings and examined relative brain size (brain to body size residuals), which, al- though a coarse measure, was available for all the studied species (21). Previously, relative brain size was found to vary with counts of field innovations (22, 23) and higher neuron numbers (24), and the size of the songbird high vocal center (HVC) vocal learning nu- cleus was found to vary positively with song repertoire size (25, 26). We found that open- ended vocal learning species had significant- ly larger relative brain sizes, but there was no significant difference in mimicking species (Fig. 4, A and B). Further, there was a signif- icant positive correlation between vocal re- pertoire size, as well as overall vocal learning complexity, and relative brain size (Fig. 4, C and D). These relationships again held when excluding vocal nonlearning and domesti- cated species (figs. S2J and S3E) or species with low sample size (vocal learning com- plexity RVLC = 0.714; PVLC = 0.0028). These findings show that vocal learning complexity, problem-solving abilities, and relatively larger brains are all related. Vocal learning cognitive relationships are not due to noncognitive factors or phylogeny We tested whether the relationships that we discovered could (i) also be found when in- cluding individual variation, (ii) be explained by noncognitive factors such as personality traits and captive conditions, or (iii) be ex- plained by phylogeny relationships. To address these factors within the same analysis, we used generalized linear mixed models of Markov chain Monte Carlo techniques (MCMCglmm). The models included phylogenetic relation- ships between species as a random effect, personality traits (shyness, or latency to feed following human disturbance; neophobia, or latency to feed in the presence of a novel ob- ject, minus shyness), experimental conditions (capture site, food deprivation period, body weight), and captive status (wild or domesti- cated) on the whole dataset of the 214 indi- vidual values. Open-endedness, mimicry, vocalization re- pertoire size, and vocal learning complexity all still significantly predicted problem-solving performance in the MCMCglmm modeling; they were still not associated with associative learning, reversal learning, or self-control, ex- cept for reversal learning, which was marginally associated with vocal learning open-endedness (Table 1 and tables S4 to S7). In addition, shy- ness was negatively associated with problem- solving and positively associated with reversal learning and self-control, neophobia was neg- atively associated with associative learning, and body weight was negatively associated with self-control but positively associated with reversal learning (Table 1 and tables S4 to S7). Although open-ended vocal learning or mimicry did not predict brain size, repertoire size and vocal learning complexity were sig- nificantly associated with relative brain size in the MCMCglmm modeling (Table 1 and tables S5 to S7). Excluding species with small sample sizes yielded similar relationships be- tween vocal learning complexity and all tested cognitive traits (table S8). Thus, MCMCglmm analyses strongly support the existence of a robust relationship between vocal learning abilities, problem-solving, and brain size, even when taking into account individual variation, phylogeny, and other potential confounding covariates. Discussion Our findings suggest a coevolution between vocal learning complexity, problem-solving, and relative brain size. Vocal learning and innovative problem-solving have separately been linked with extinction risk, fitness, and Audet et al., Science 381, 1170–1175 (2023) 15 September 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E sexual selection (19, 27). Further, both problem- solving and bird songs vary according to habitat differences, for example, urbanization (28, 29). To explain our and these ecological findings, we suggest that a selective factor links these traits within a “vocal learning cognitive com- plex.” This selective factor could be a genetic component that drives coevolution of vocal learning complexity, problem-solving, and relative brain size. Previous studies that performed within- species comparisons on vocal learning and other cognitive traits found no or conflicting evidence of relationships [(6, 7, 9, 10, 12), re- viewed in (5)]. A major factor as to why we find a clear relationship between vocal learn- ing complexity and problem-solving is likely greater differences across than within spe- cies. In addition, some studies trained birds to first solve problems and then measured their capacity to repeat the learned solution, whereas we measured problem-solving on the first trial. We also generated a new measure of vocal learning complexity that did not ig- nore calls that were assumed to be innate but included calls and songs together. Calls are increasingly recognized as learned in vocal learning species (30, 31). Our finding of a link between vocal learning complexity and brain size is consistent with the prior finding that two of the three avian vocal learning lineages (songbirds and parrots) have larger relative brain sizes and a higher density of telencephalic neurons compared with vocal nonlearners (32). Devoogd et al. (25) found a relationship between song rep- ertoire size and the HVC size across songbird species, but not telencephalon size [confirmed in (26)]. We hypothesize that the relative in- crease in brain size across species could be due to relative increases in the song system and the adjacent nonvocal motor circuit (33) that is potentially involved in problem-solving. Vocal learners can also synchronize body move- ments to rhythmic sounds of music (dance), which is thought to be controlled by the sur- rounding motor circuits (34, 35) and could be another component of a vocal learning cog- nitive complex. Another brain region to con- sider is the caudal-lateral nidopallium (NCL), or avian prefrontal cortex, which is involved in complex cognitive processing (36). Regard- less of specific brain regions, a higher density of neurons likely provides vocal learners with more brain computational power. At the molecular level, expression levels of different N-methyl-D-aspartate (NMDA) glu- tamate receptor subunits in song nuclei and nearby brain regions have been associated with both complex vocal learning capacity (37, 38) and problem-solving (39). Thus, this neuro- transmitter receptor family is a plausible can- didate to partly explain our discovered cognitive relationships. There has been a long-standing assumption of a link between human spoken language, ad- vanced cognition, and larger brain size relative to other species (40). Our study quantitative- ly tests such a relationship between species in a vocal learning bird lineage and serves as a model for testing other avian and mamma- lian lineages. Testing cognitive abilities across more-divergent species might be more chal- lenging because it may require different ap- paratus designs, whereas this is not the case for comparative genomic studies across spe- cies without behavioral testing [e.g., (41)]. For example, some bird lineages have highly diver- gent beak shapes (e.g., hummingbirds, pelicans) or rely mainly on their feet to manipulate ob- jects (e.g., parrots, birds of prey). Moreover, the vocal repertoires of many nonsongbird spe- cies are not as well characterized as those of songbirds. Nevertheless, it would be interest- ing to see whether the relationships we discov- ered here exist for other vocal learning species and innate repertoires of vocal nonlearning species. Our results support the continuum hypothesis of vocal learning within a clade (1). More broadly, our discoveries open the door for an unexplored sphere of research on shared neurobiological, molecular, and physiological evolutionary foundations of vocal learning and problem-solving. RE FERENCES AND NOTES 1. E. D. Jarvis, Science 366, 50–54 (2019). 2. V. M. Janik, P. J. B. Slater, Anim. Behav. 60, 1–11 3. (2000). J. O. Van Horik, N. S. Clayton, N. J. Emery, in The Oxford Handbook of Comparative Evolutionary Psychology, T. K. Shackelford, J. Vonk, Eds. (Oxford Univ. Press, 2012). 4. S. M. Reader, K. N. Laland, Eds., Animal Innovation (Oxford Univ. Press, ed. 1, 2003). 5. W. A. Searcy, S. Nowicki, Anim. Behav. 151, 217–227 (2019). 6. N. J. Boogert, L.-A. Giraldeau, L. Lefebvre, Anim. Behav. 76, 1735–1741 (2008). 7. N. J. Boogert, R. C. Anderson, S. Peters, W. A. Searcy, S. Nowicki, Anim. Behav. 81, 1209–1216 (2011). 8. C. N. Templeton, K. N. Laland, N. J. Boogert, Anim. Behav. 92, 63–71 (2014). 9. R. C. Anderson et al., Anim. Cogn. 20, 309–320 (2017). 10. A. L. DuBois, S. Nowicki, S. Peters, K. D. Rivera-Cáceres, W. A. Searcy, Anim. Behav. 137, 205–213 (2018). 11. K. B. Sewall, J. A. Soha, S. Peters, S. Nowicki, Biol. Lett. 9, 20130344 (2013). 12. T. M. Farrell, K. Weaver, Y.-S. An, S. A. MacDougall-Shackleton, Behav. Ecol. 23, 101–111 (2012). 13. K. L. Schmidt, S. D. Moore, E. A. MacDougall-Shackleton, S. A. MacDougall-Shackleton, Anim. Behav. 86, 25–35 (2013). 14. S. C. Vernes et al., Philos. Trans. R. Soc. Lond. Ser. B 376, 20200236 (2021). 15. M. Wirthlin et al., Neuron 104, 87–99 (2019). 16. W. A. Searcy, J. Soha, S. Peters, S. Nowicki, Philos. Trans. R. Soc. Lond. Ser. B 376, 20200257 (2021). 17. P. Rundstrom, N. Creanza, Curr. Opin. Neurobiol. 67, 228–239 (2021). 18. C. M. Robinson, K. T. Snyder, N. Creanza, eLife 8, e44454 (2019). 19. J.-N. Audet, Integr. Comp. Biol. 60, 955–966 (2020). 20. J.-N. Audet, L. Lefebvre, Behav. Ecol. 28, 943–947 (2017). 21. F. Sayol, D. Sol, A. L. Pigot, Front. Ecol. Evol. 8, 58 (2020). 22. L. Lefebvre, S. M. Reader, D. Sol, Brain Behav. Evol. 63, 233–246 (2004). 23. S. Benson-Amram, B. Dantzer, G. Stricker, E. M. Swanson, K. E. Holekamp, Proc. Natl. Acad. Sci. U.S.A. 113, 2532–2537 (2016). 24. D. Sol et al., Nat. Ecol. Evol. 6, 1381–1389 (2022). 25. T. J. Devoogd, J. R. Krebs, S. D. Healy, A. Purvis, Proc. Biol. Sci. 254, 75–82 (1993). 26. L. Z. Garamszegi, M. Eens, J. Erritzøe, A. P. Møller, Behav. Ecol. 16, 335–345 (2005). 27. S. A. Macdougall-Shackleton, in Current Ornithology, V. Nolan, E. D. Ketterson, C. F. Thompson, Eds. (Springer, 1997), pp. 81–124. 28. J.-N. Audet, S. Ducatez, L. Lefebvre, Behav. Ecol. 27, 637–644 (2016). 29. J. L. Dowling, D. A. Luther, P. P. Marra, Behav. Ecol. 23, 201–209 (2012). 30. P. Marler, Ann. N. Y. Acad. Sci. 1016, 31–44 (2004). 31. K. B. Sewall, A. M. Young, T. F. Wright, Anim. Behav. 120, 163–172 (2016). 32. S. Olkowicz et al., Proc. Natl. Acad. Sci. U.S.A. 113, 7255–7260 (2016). 33. G. Feenders et al., PLOS ONE 3, e1768 (2008). 34. A. D. Patel, Philos. Trans. R. Soc. Lond. Ser. B 376, 20200326 (2021). 35. M. Chakraborty, E. D. Jarvis, Philos. Trans. R. Soc. Lond. Ser. B 370, 20150056 (2015). 36. O. Güntürkün, T. Bugnyar, Trends Cogn. Sci. 20, 291–303 (2016). 37. F. S. Livingston, R. Mooney, J. Neurosci. 17, 8997–9009 (1997). 38. M. E. Basham, E. J. Nordeen, K. W. Nordeen, Neurobiol. Learn. Mem. 66, 295–304 (1996). 39. J.-N. Audet et al., Sci. Adv. 4, eaao6369 (2018). 40. T. W. Deacon, Symbolic Species (Norton, 1998). 41. Y. Toda et al., Science 373, 226–231 (2021). 42. C. R. Cooney et al., Nature 542, 344–347 (2017). 43. F. Sayol, P. A. Downing, A. N. Iwaniuk, J. Maspons, D. Sol, Nat. Commun. 9, 2820 (2018). 44. J.-N. Audet, M. Couture, E. D. Jarvis, Songbird species that display more-complex vocal learning are better problem solvers and have larger brains. Dryad (2023); https://doi.org/ 10.5061/dryad.tb2rbp06n. AC KNOWLED GME NTS We thank G. Smith-Vidaurre for insightful discussions on vocal learning complexity; S. Ducatez for assisting with the statistical analyses; L. Lefebvre, M. Davenport, and L. Cauchard for providing helpful suggestions on the manuscript; C. Scivolette for helping with bird captures; S. Sepe for helping with the establishment of wild-bird care protocols; G. Permuy and L. Tchernichovski for helping with bird care; and W. T. Gale for assisting with the construction of the behavior laboratory at the Rockefeller University Field Research Center. Funding: This study was supported by a Banting postdoctoral fellowship and an NSERC postdoctoral fellowship to J.-N.A. and by the Howard Hughes Medical Institute and a Keck Foundation Award to E.D.J. Author contributions: Conceptualization: J.-N.A., M.C., E.D.J.; Methodology: J.-N.A., M.C.; Investigation: J.-N.A., M.C.; Visualization: J.-N.A., M.C.; Funding acquisition: J.-N.A., E.D.J.; Supervision: J.-N.A., E.D.J.; Writing – original draft: J.-N.A.; Writing – review and editing: J.-N.A., M.C., E.D.J. Competing interests: The authors declare that they have no competing interests. Data and materials availability: The raw dataset and code are available at Dryad (44). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh3428 Materials and Methods Figs. S1 to S3 Tables S1 to S8 References (45–144) MDAR Reproducibility Checklist Movies S1 to S7 Submitted 25 February 2023; accepted 14 August 2023 10.1126/science.adh3428 Audet et al., Science 381, 1170–1175 (2023) 15 September 2023 6 of 6
10.1126_science.adh2526
Corrected 5 January 2024. See full text. RES EARCH ATMOSPHERES Iodine oxoacids enhance nucleation of sulfuric acid particles in the atmosphere Xu-Cheng He1,2,3*, Mario Simon4, Siddharth Iyer5, Hong-Bin Xie6*, Birte Rörup1, Jiali Shen1, Henning Finkenzeller7,8, Dominik Stolzenburg1,9, Rongjie Zhang6, Andrea Baccarini10,11, Yee Jun Tham1,12, Mingyi Wang13, Stavros Amanatidis13, Ana A. Piedehierro3, Antonio Amorim14, Rima Baalbaki1, Zoé Brasseur1, Lucía Caudillo4, Biwu Chu1,15, Lubna Dada1,10, Jonathan Duplissy1,16, Imad El Haddad10, Richard C. Flagan13, Manuel Granzin4, Armin Hansel17, Martin Heinritzi4, Victoria Hofbauer2, Tuija Jokinen1,18, Deniz Kemppainen1, Weimeng Kong13, Jordan Krechmer19, Andreas Kürten4, Houssni Lamkaddam10, Brandon Lopez2,20, Fangfang Ma6, Naser G. A. Mahfouz2, Vladimir Makhmutov21,22, Hanna E. Manninen23, Guillaume Marie4, Ruby Marten10, Dario Massabò24, Roy L. Mauldin25,26, Bernhard Mentler17, Antti Onnela23, Tuukka Petäjä1, Joschka Pfeifer23, Maxim Philippov21, Ananth Ranjithkumar27, Matti P. Rissanen5,28, Siegfried Schobesberger29, Wiebke Scholz17, Benjamin Schulze13, Mihnea Surdu10, Roseline C. Thakur1, António Tomé30, Andrea C. Wagner4, Dongyu Wang10, Yonghong Wang1,15, Stefan K. Weber23,4, André Welti3, Paul M. Winkler31, Marcel Zauner-Wieczorek4, Urs Baltensperger10, Joachim Curtius4, Theo Kurtén28, Douglas R. Worsnop1,19, Rainer Volkamer7,8, Katrianne Lehtipalo1,3, Jasper Kirkby23,4, Neil M. Donahue2,20,25,32, Mikko Sipilä1*, Markku Kulmala1,16,33,34* The main nucleating vapor in the atmosphere is thought to be sulfuric acid (H2SO4), stabilized by ammonia (NH3). However, in marine and polar regions, NH3 is generally low, and H2SO4 is frequently found together with iodine oxoacids [HIOx, i.e., iodic acid (HIO3) and iodous acid (HIO2)]. In experiments performed with the CERN CLOUD (Cosmics Leaving OUtdoor Droplets) chamber, we investigated the interplay of H2SO4 and HIOx during atmospheric particle nucleation. We found that HIOx greatly enhances H2SO4(-NH3) nucleation through two different interactions. First, HIO3 strongly binds with H2SO4 in charged clusters so they drive particle nucleation synergistically. Second, HIO2 substitutes for NH3, forming strongly bound H2SO4-HIO2 acid-base pairs in molecular clusters. Global observations imply that HIOx is enhancing H2SO4(-NH3) nucleation rates 10- to 10,000-fold in marine and polar regions. A erosols influence climate by acting as cloud condensation nuclei (CCN) and by scattering solar radiation. Secondary aerosol and CCN formation continue to be two of the largest uncertainties hin- dering accurate projection of climate change (1). Only a few types of vapors in the atmosphere can nucleate to form new aerosol particles, which can further grow to CCN sizes. Sulfuric acid (H2SO4) is considered to be the primary vapor (2) driving particle formation in the atmo- sphere of both polluted environments (3, 4) and pristine environments (5–7). However, as H2SO4-H2O binary nucleation is slow, stabilizing vapors, such as ammonia (NH3), amines, and oxidized organics, are generally needed to explain observed particle formation rates (3–11). In terms of radiative balance, marine clouds, especially low-level marine stratocumulus (12), are key players because they have strong long- wave emission and efficiently reflect solar radia- tion back to space. As marine cloud formation is often limited by low CCN number concen- trations, it is important to reach a comprehen- sive understanding of new particle formation in marine environments. New particle and subsequent CCN formation in marine regions is presently thought to be driven by H2SO4 and methanesulfonic acid (MSA) (8, 13), aided by NH3 (5, 14). However, a recent global survey of aerosol acidity suggests that global models substantially overestimate NH3 concentrations; in particular, the polar atmosphere and high altitudes are characterized by low NH3 con- centrations (15). Assuming solely H2SO4 nucle- ation, advanced Earth system models struggle to reproduce aerosol number concentrations measured by aircraft (16), leading to low con- fidence for estimates of aerosol radiative forc- ing. Iodine-driven nucleation (17–21) has not yet been incorporated into Earth system mod- els; iodine oxoacids (HIOx, x = 2 to 3 in this study) can drive rapid particle formation under low NH3 conditions, and they may play an im- portant role in polar, marine, and free tropo- spheric particle formation. In the marine atmosphere, iodine and sulfur precursors emitted from the ocean surface lead to the formation of both H2SO4 and HIOx (22). HIOx has generally been observed at con- centrations similar to or lower than H2SO4 (6, 18, 21, 23). Despite the higher nucleation potential of HIOx compared with H2SO4 (18), iodine-driven new particle formation has hith- erto been considered important only in re- gions with considerably higher concentrations of iodic acid (HIO3) than of H2SO4, such as coastal zones and specific regions in the Arctic (17, 18, 20, 21, 24, 25). However, new particle formation from the mixed chemical system HIOx-H2SO4(-NH3) has not been reported so far. Particle formation experiments in CLOUD Here we report laboratory experiments per- formed in the CERN CLOUD (Cosmics Leaving OUtdoor Droplets) chamber (5) (see methods in the supplementary materials for details) between September 2018 and December 2019 under conditions relevant for marine and polar environments. We performed particle forma- tion experiments using HIOx-H2SO4(-NH3) va- pors produced from the following precursors: molecular iodine (I2), sulfur dioxide (SO2), ammo- nia (NH3), ozone (O3), and water vapor (H2O). 1Institute for Atmospheric and Earth System Research/Physics, Faculty of Science, University of Helsinki, 00014 Helsinki, Finland. 2Center for Atmospheric Particle Studies, Carnegie Mellon University, Pittsburgh, PA 15213, USA. 3Finnish Meteorological Institute, 00560 Helsinki, Finland. 4Institute for Atmospheric and Environmental Sciences, Goethe University Frankfurt, 60438 Frankfurt am Main, Germany. 5Aerosol Physics Laboratory, Faculty of Engineering and Natural Sciences, Tampere University, 33720 Tampere, Finland. 6Key Laboratory of Industrial Ecology and Environmental Engineering (Ministry of Education), School of Environmental Science and Technology, Dalian University of Technology, 116024 Dalian, China. 7Department of Chemistry, University of Colorado Boulder, Boulder, CO 80309, USA. 8Cooperative Institute for Research in Environmental Sciences, University of Colorado Boulder, Boulder, CO 80309, USA. 9Institute for Materials Chemistry, TU Wien, 1060 Vienna, Austria. 10Laboratory of Atmospheric Chemistry, Paul Scherrer Institute, CH-5232 Villigen, Switzerland. 11Laboratory of Atmospheric Processes and their Impact, Ecole Polytechnique Fédérale de Lausanne (EPFL), 1015 Lausanne, Switzerland. 12School of Marine Sciences, Sun Yat-sen University, 519082 Zhuhai, China. 13Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, CA 91125, USA. 14CENTRA and Faculdade de Ciências da Universidade de Lisboa, 1749-016 Lisboa, Portugal. 15Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, 100084 Beijing, China. 16Helsinki Institute of Physics, University of Helsinki, 00014 Helsinki, Finland. 17Institute of Ion Physics and Applied Physics, University of Innsbruck, 6020 Innsbruck, Austria. 18Climate and Atmosphere Research Centre (CARE-C), The Cyprus Institute, 1645 Nicosia, Cyprus. 19Aerodyne Research, Inc., Billerica, MA 01821, USA. 20Department of Chemical Engineering, Carnegie Mellon University, Pittsburgh, PA 15213, USA. 21P. N. Lebedev Physical Institute of the Russian Academy of Sciences, 119991 Moscow, Russia. 22Moscow Institute of Physics and Technology (National Research University),141701 Moscow, Russian Federation. 23CERN, the European Organization for Nuclear Research, CH-1211 Geneva, Switzerland. 24Department of Physics, University of Genoa, 16146 Genoa, Italy. 25Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213, USA. 26Department of Atmospheric and Oceanic Sciences, University of Colorado Boulder, Boulder, CO 80309, USA. 27Natural Environment Research Council, British Antarctic Survey, CB3 0ET Cambridge, UK. 28Department of Chemistry, University of Helsinki, 00014 Helsinki, Finland. 29Department of Applied Physics, University of Eastern Finland, 70211 Kuopio, Finland. 30Instituto Dom Luiz (IDL)–Universidade da Beira Interior, 6201-001 Covilhã, Portugal. 31Faculty of Physics, University of Vienna, 1090 Wien, Austria. 32Department of Engineering and Public Policy, Carnegie Mellon University, Pittsburgh, PA 15213, USA. 33Joint International Research Laboratory of Atmospheric and Earth System Sciences, School of Atmospheric Sciences, Nanjing University, 210023 Nanjing, China. 34Aerosol and Haze Laboratory, Beijing Advanced Innovation Center for Soft Matter Science and Engineering, Beijing University of Chemical Technology, 100029 Beijing, China. *Corresponding author. Email: xucheng.he@helsinki.fi (X.-C.H.); markku.kulmala@helsinki.fi (M.K.); mikko.sipila@helsinki.fi (M.Sip.); hbxie@dlut.edu.cn (H.-B.X.) He et al., Science 382, 1308–1314 (2023) 15 December 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Corrected 5 January 2024. See full text. A a b NH3 detection limit B b NH 3 concentration H2SO4 concentration HIO 3 concentration HIO 2 concentration d c e f g C D 9 10 8 10 7 10 6 10 5 10 4 10 10 1 0.1 0.01 0.001 ) 3 - m c ( n o i t a r t n e c n o C ) 1 - s 3 - m c ( 7 . 1 J , e t a r n o i t a e c u N l 0.0001 00:00 Oct-10,2018 Measured J 1.7 Expected J 1.7 from H 2SO4 + NH 3 01:00 02:00 03:00 04:00 05:00 22:00 23:00 Time (UTC) 00:00 Oct-11,2018 01:00 02:00 9 10 8 10 7 10 6 10 5 10 4 10 10 1 0.1 0.01 0.001 0.0001 Fig. 1. New particle formation from HIOx-H2SO4 and HIOx-H2SO4-NH3 at −10°C. (A and B) vapor concentrations and (C and D) nucleation rates. Solid black lines show the measured nucleation rates at 1.7 nm and solid red lines present predicted J1.7 from H2SO4-NH3 nucleation alone (14). Dashed lines represent vapor concentrations, and vertical gray bars show experimental stages. The experiments show that the rapid nucleation rates cannot be explained by the H2SO4-NH3 mechanism alone. HIOx significantly enhances H2SO4-NH3 nucleation at comparable HIO3 and H2SO4 concentrations. The NH3 concentra- tion in (A) is below the detection limit of the H3O+-CIMS (~4 pptv). An NH3 concentration of 4 pptv is used to conservatively estimate the H2SO4-NH3 nucleation rates in (C). The experimental conditions are 41.1 parts per billion by volume (ppbv) O3, 63.5% relative humidity (RH), 2.3 ppbv SO2, and 17.4 pptv I2 [(A) and (C)]; and 40.8 ppbv O3, 62.3% RH, 1.6 ppbv SO2, and 67.2 pptv I2 [(B) and (D)]. Stages a, c, d, e, f, and g enhanced the UVH light intensity (higher OH production rates), and stage b increased the green light intensity (higher I2 photolysis rate). To investigate possible synergies in HIOx- H2SO4(-NH3) nucleation, green and ultraviolet light sources were used to drive photochem- ical production of HIOx and H2SO4 from I2 and SO2. An example experiment at −10°C is shown in Fig. 1 and fig. S1, and at 10°C in fig. S2. Experiments were first performed with- out any added NH3 [<4 parts per trillion by volume (pptv) contaminant level]; these are shown in the left-hand panels of Fig. 1 and figs. S1 and S2. A second set of experiments were performed with NH3 added to the cham- ber (right-hand panels of Fig. 1 and figs. S1 and S2). At both temperatures, a significantly higher nucleation rate at 1.7 nm, J1.7, is observed in the presence of HIOx than the J1.7 expected from H2SO4-NH3 nucleation (5, 14), both with- out and with added NH3. In Fig. 2, we present J1.7 for the HIOx-H2SO4 system (hollow markers) and the HIOx-H2SO4- NH3 system (filled markers) at 10°C (circles) and −10°C (squares). The concentration ranges of HIOx and H2SO4 closely match ambient val- ues, spanning from <106 cm−3 to nearly 108 cm−3 (6, 17, 18, 20, 21, 23). We show the measured J1.7 for these mixed systems for various possible drivers: H2SO4 (Fig. 2A), HIO3 + H2SO4 (Fig. 2B), and (HIO3 + H2SO4) × HIO2 (Fig. 2C) (HIO2, iodous acid). The data at both temperatures be- come progressively less scattered when plotted against these variables, as well as more con- sistent with parameterizations (14, 18). The H2SO4-NH3 mechanism cannot predict the nu- cleation rates, even when the HIOx concentra- tion is much lower than that of H2SO4 (Fig. 2A). For instance, J1.7 at 10°C from HIOx-H2SO4 with NH3 < 4 pptv (Fig. 2A, hollow circles) is roughly 60 times faster than J1.7 from H2SO4 with NH3 at 4 pptv; this is as fast as nucleation from H2SO4 with NH3 at 500 pptv. Therefore, sub-pptv levels of HIOx are as effective at stabilizing H2SO4 as 500 pptv of NH3. Hence, HIOx may replace NH3 as a nucleation driver in pristine marine and polar environments, where NH3 concentrations are typically below a few tens of parts per trillion by volume or lower (26, 27). Figure 2B shows the observed J1.7 versus total acid concentration (HIO3 + H2SO4) and compares these rates to the values predicted by the H2SO4(-NH3) parameterizations (14), applying (HIO3 + H2SO4) as H2SO4. The J1.7 of the HIOx-H2SO4 system without added NH3 (hollow markers) remains higher than the pre- diction for H2SO4(-NH3) nucleation. This indi- cates that HIOx contributes more prominently to nucleation than by simply increasing the acid concentration. Moreover, the relatively mild sensitivity to NH3 suggests that the base sta- bilization comes from another source. This is supported by Fig. 2C, which indicates that HIO2 is effectively providing base stabilization in the molecular clusters. To further investi- gate the underlying mechanisms, we studied the molecular composition of nucleating parti- cles under neutral (ion-free) and charged (ion- induced) conditions, as described below. HIO2 accelerates neutral nucleation To measure neutral clusters, we used a nitrate chemical ionization mass spectrometer (nitrate- CIMS). The concentrations of monomers HIO3, H2SO4, and HIO2 are presented in Fig. 3A, together with four product dimers in Fig. 3B. Although the HIO2 concentration was one to two orders of magnitude lower than that of HIO3 or H2SO4, the most prominent dimers, HIO3-HIO2 and H2SO4-HIO2, both contain HIO2. He et al., Science 382, 1308–1314 (2023) 15 December 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Corrected 5 January 2024. See full text. A 4 2 100 6 4 ) 1 - s 3 - m c ( 7 . 1 J 2 10 6 4 , e t a r n o i t a e c u N l 2 1 6 4 2 0.1 6 4 2 0.01 4 6 8 106 2 4 6 8 107 H2SO4 (cm-3) B HIO3 4 2 8 6 4 2 7 10 6 10 2 4 6 8 108 2 C –10°C 10°C HIOx H2SO4+NH3(500ppt) H2SO4+NH3(4ppt) HIOx+H2SO4 HIOx+H2SO4+NH3 ACDC: HIOx+H2SO4(Neutral) ACDC: HIO3+HIO2 or H2SO4+HIO2(Neutral) 4 6 8 108 1011 2 4 6 1012 2 4 6 1013 2 4 6 2 1014 (HIO3 + H2SO4) HIO2 (cm-6) 2 4 6 8 107 HIO3 + H2SO4 (cm-3) Fig. 2. Nucleation rates of HIOx-H2SO4(-NH3) systems. Nucleation rates at 1.7 nm versus (A) H2SO4, (B) HIO3 + H2SO4, and (C) (HIO3 + H2SO4) × HIO2 at +10° and −10°C. All data points and lines show experiments carried out at galactic cosmic ray ionization conditions, except for the atmospheric cluster dynamics code (ACDC) simulations in (C) (orange band and filled diamonds), which represent the theoretical prediction for the neutral nucleation rates (see methods). The color bar represents HIO3 concentration (per cubic centimeter). H2SO4-NH3 mechanism fails to predict the overall nucleation rates, even with HIOx is much lower than H2SO4. The J1.7 from experiments with high H2SO4 is also higher than that predicted by pure iodine oxoacids (18). The nucleation rates become less spread when plotted against (HIO3 + H2SO4) × HIO2, as well as more consistent with parameterizations and ACDC predictions. The results show that HIO3 and HIO2 have to be considered together with H2SO4 to predict the nucleation rates in this multicomponent system. H2SO4-NH3 nucleation rates (dotted and dash-dotted lines) are calculated following Dunne et al. (14), whereas HIOx nucleation rates (solid lines) are calculated on the basis of J1.7, HIO3, and recalculated HIO2 from He et al. (18), applying HIO3 × HIO2 as (HIO3 + H2SO4) × HIO2, to guide the eye. The experimental conditions for HIOx-H2SO4(-NH3) experiments are 38.4 to 53.2 ppbv O3, 41.9 to 75.3% RH, 0.6 to 11.2 ppbv SO2, and 10.0 to 57.7 pptv I2. The NH3 concentrations for the filled squares and filled circles range from 30 to 42 pptv and from 176 to 261 pptv, respectively. The error bars show one standard deviation during the data selection periods. Overall systematic scale errors on the HIO3 concentrations of −33% and +50% and on the nucleation rates of a factor of 10 are not shown on the data points. Despite HIO3-HIO2 clusters having been re- ported before (18, 28), we believe this is the first observation of H2SO4-HIO2 dimers. While HIO2 enables H2SO4-HIO2 dimer for- mation, its role in larger clusters is not clear. We address this with a combination of quan- tum chemical calculations and cluster dynam- ics modeling (29). We optimized the geometries of H2SO4-HIO2, HIO3-HIO2, and H2SO4-HIO3- HIO2 clusters and calculated their formation free energies and evaporation rates (fig. S3). Clusters containing HIO2 are the most stable and, moreover, show an exceptionally wide range of stable combinations of molecules. The cluster geometries suggest that HIO2 en- hances H2SO4 neutral nucleation in the same way as it does for HIO3 neutral nucleation (18). Specifically, HIO2 accepts the proton donated either by H2SO4 or HIO3, thereby function- ing as a base. Furthermore, HIO2 forms strong halogen bonds with H2SO4 and HIO3, further enhancing the cluster stability. Clusters includ- ing HIO2 are even more stable than H2SO4- DMA (dimethyl amine) clusters (fig. S3), which is known to cluster at the collision limit for sulfuric acid with only 4 pptv DMA (3). How- ever, the predicted neutral nucleation rates for the H2SO4-HIO2 and HIO3-HIO2 systems still underestimate our measured nucleation rates [galactic cosmic ray (GCR) conditions, the sum of neutral and ion-induced chan- nels] at −10°C (Fig. 2C, orange band). On the other hand, the predicted HIOx-H2SO4 neu- tral nucleation rates approximately agree with CLOUD observations (Fig. 2C, squares and di- amonds). The consistency between theoretical predictions and the CLOUD measurements at −10°C suggests that neutral nucleation domi- nates at this temperature, which is also indicated by the fact that the nucleation rates far exceed the ion-pair production rate limit (2 to 10 cm−3 s−1). Additionally, this suggests that the control- ling mechanism is indeed a synergy of three molecules (HIO3, H2SO4, and HIO2) and not simply the combined neutral nucleation of any two molecules. Given that HIO2 behaves as a base, we show in Fig. 2C our observed J1.7 versus (HIO3 + H2SO4) × HIO2. This expres- sion is proportional to the formation rate of the dimer (H2SO4-HIO2 or HIO3-HIO2), which rep- resents the initial nucleating cluster. We find that the HIOx-H2SO4(-NH3) nucleation rates fall near the prediction from HIOx nucleation (J1.7 versus HIO3 × HIO2; H2SO4 is absent in pure iodine oxoacid nucleation, but it is added to the HIO3 concentration given its identical role) (18), implying that HIO2 indeed plays the key role as stabilizer both for HIO3 and H2SO4 and that NH3 plays a minor role. While the formation mechanism for HIO3 has recently been established (22), the path- way for HIO2 formation remains uncertain. A quantum chemical study provided a potential energy surface describing formation of HIO2 from iodooxy hypoiodite, I2O2 + H2O (30). We have extended this study with high-level quan- tum chemical calculations and provide a re- vised potential energy surface in fig. S4A. We also present a potential new pathway for HIO2 formation from iodine dioxide (OIO) and the hydroperoxyl radical (HO2) (fig. S4B). Our cal- culations show that both the singlet and trip- let channels are exothermic. Further studies are needed to quantify the relative importance in the atmosphere of these two channels. Because complex reactions are involved in the formation of HIO3 and HIO2, it is impor- tant to confirm that the HIO3:HIO2 ratio in the CLOUD chamber matches ambient conditions. Figure S5 shows that both the ratio and ab- solute concentrations of HIO3 and HIO2 fall within the range measured at Mace Head (17) and Ny-Ålesund (21), confirming that the results from our study are relevant to the atmosphere. HIO3 enhances ion-induced nucleation Ions can stabilize embryonic molecular clus- ters, leading to ion-induced nucleation (IIN) (5, 18, 19, 31, 32). To investigate the influence of ions on HIOx-H2SO4 nucleation, we increased the ionization rate in the chamber in three steps at 10°C: (i) neutral (ion-free), (ii) GCR ion- ization (∼1000 ion pairs cm−3), and (iii) beam- enhanced ionization (∼6000 ion pairs cm−3) (fig. S6). Compared with neutral conditions, J1.7 at GCR ionization rates is enhanced by He et al., Science 382, 1308–1314 (2023) 15 December 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Corrected 5 January 2024. See full text. Fig. 3. Neutral and charged cluster composition during HIOx-H2SO4- NH3 nucleation. Background- subtracted neutral monomer (A) and dimer (B) concentrations in HIOx- H2SO4 nucleation events at +10°C (pink bars) and −10°C (cyan bars). (C and D) Negatively charged cluster compositions of HIOx-H2SO4 nucleation at +10° and −10°C, respectively. (E and F) Negatively charged cluster compositions of HIOx-H2SO4-NH3 nucle- ation at +10° and −10°C, respectively. As indicated in (B), the dominant neutral dimers are H2SO4-HIO2 and HIO3-HIO2 clusters—despite very low HIO2 concentrations—which represent the initial molecular clusters during neutral nucleation. Ion-induced nucleation is dominated by charged HIO3-H2SO4 [(C) and (D)] or HIO3-H2SO4(-NH3) [(E) and (F)] cluster formation processes. HIO3-NH3 clusters are not detected, which suggests that NH3 has a negligible effect on ion-induced HIO3 cluster formation. The marker size is shown in the legend (cps, ion counts per second). The experimental conditions are 38.5 to 43.9 ppbv O3, 61.6 to 75.2% RH, 0.7 to 11.0 ppbv SO2, and 14.4 to 44.5 pptv I2. ) 3 - m c ( n o i t a r t n e c n o C 0.2 0.0 -0.2 -0.4 -0.6 0.2 0.0 -0.2 -0.4 -0.6 ) h T ( t c e f e d s s a M A neutral monomer B neutral dimer +10 °C 10 °C 107 106 105 104 H2SO4 HIO3 HIO2 H2SO4·H2SO4 H2SO4·HIO3 H2SO4·HIO2 HIO3·HIO2 C HIOx + H2SO4 at +10 °C D HIO x + H2SO4 at 10 °C < 0.01 cps 0.01-0.1 0.1-1 1-10 10 E HIO x + H2SO4 + NH3 at +10 °C F HIOx + H2SO4 + NH3 at 10 °C Other HIOx and iodine containing H2SO4 and sulfur containing HIOx - H2SO4 H2SO4 - NH3 HIOx - H2SO4 - NH3 0 200 400 600 800 1000 0 200 Mass/charge, m/z (Th) 400 600 800 1000 ∼50 times at 2 × 107 cm−3 H2SO4 and 5 × 106 cm−3 HIO3. As with HIOx, ion-induced HIOx-H2SO4 nucleation only occurs with negative ions (com- pare fig. S6A and fig. S6B). Interestingly, six times larger ion concentrations formed by the pion beam only enhance J1.7 by a factor of two. This is likely because the increased ion- ion recombination rate, and hence the shorter charge lifetime, neutralizes some clusters be- fore they have become stable against evapo- ration when neutral. When NH3 is added to the HIOx-H2SO4 system, it initiates positive IIN that is as strong as negative IIN at −10°C (fig. S1D). Adding NH3 approximately doubles the overall J1.7. To measure the molecular composition of charged clusters, we used an atmospheric pres- sure interface time-of-flight mass spectrome- ter. For HIOx-H2SO4 IIN without NH3 injection (Fig. 3, C and D), we observe a series of charged clusters with the empirical formula (HIO3)n- − (cyan triangles), which indicate (H2SO4)m-HSO4 synergistic IIN of HIO3 and H2SO4. We identify these clusters as (n+m+1)-mer (which include the ion; fig. S7). At 10°C, monomers, dimers, and trimers consist primarily of H2SO4, whereas HIO3 appears in clusters starting from the tetra- mers and becomes equal to the H2SO4 mole fraction already in the hexamers. At −10°C, HIO3 appears in the dimers and becomes equal to the H2SO4 starting with the tetramers. In these experiments, the HIO3:H2SO4 ratio in the gas phase is between 0.3 and 1.4, and the molar ratios of I:S in the larger clusters tend toward 1:1. We also know that pure ion-induced HIO3 nucleation proceeds at the collision limit (18, 19) but that ion-induced H2SO4 nucleation is slower than the collision limit (5). We there- fore conclude that H2SO4 condensation is en- hanced by HIO3 for a cluster stoichiometry up to 1:1, beyond which the net rate of H2SO4 con- densation slows, while HIO3 condensation is limited by the collision rate under our exper- imental conditions. We performed additional experiments in which NH3 was added to the HIOx-H2SO4 sys- tem. Notably, none of the charged pure iodine (i.e., H2SO4-free) clusters contained NH3, which indicates a negligible role of NH3 in ion-induced HIO3 nucleation. This was independently con- firmed by raising NH3 from the background level (<4 pptv) to 100 pptv in an iodine oxoacid nucleation experiment without H2SO4 (fig. S8). The measured nucleation rate at 1.7 nm remained constant throughout the experi- ment, indicating that HIO3(-HIO2) nucleation is unaffected by NH3. On the other hand, we found a set of clusters − with the composition (H2SO4)n-(NH3)m-HSO4 − in the and (HIO3)n-(H2SO4)m-(NH3)j-HSO4 mass spectra of charged clusters (Fig. 3, E and F), similar to the clusters reported near the coast of Antarctica (6). We found that NH3 is only present in charged tetramers and above, consistent with its behavior in H2SO4-NH3 IIN (5). The iodine and sulfur molar fraction dis- tributions remained unchanged after adding NH3 to the system, likely because the HIO3- H2SO4 negative IIN had already reached the collision limit (fig. S7). The presence of NH3 only converted some of the (HIO3)n-(H2SO4)m- − ions − to (HIO3)n-(H2SO4)m-(NH3)j-HSO4 HSO4 and gives rise to positive IIN (fig. S1). Particle growth Since the atmospheric concentration of HIO2 is less than one-tenth that of HIO3, its role in particle growth is minor (18). To evaluate the role of HIO3 and H2SO4 in particle growth, we compare in fig. S9A our measured growth rates between 1.8 and 3.2 nm (GR1.8-3.2) with those calculated assuming condensation of H2SO4 and HIO3 (18, 33) at the collision limit. The good agreement indicates that H2SO4 and HIO3 are the main condensing vapors driving particle growth (18, 33) while other iodine species contribute little to particle mass. We show in fig. S9B the measured and predicted particle survival probability, J2.5/J1.7, which increases at faster growth rates and approaches unity above growth rates of ~10 nm hour−1 for the CLOUD chamber (2.2 × 10−3 s−1 wall loss rate). In the marine atmosphere, condensation He et al., Science 382, 1308–1314 (2023) 15 December 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Corrected 5 January 2024. See full text. 78-90N A Arctic Ocean: Aug-Sep Villum: March-Aug Ny-Ålesund: April-Aug 81N 79N 70N 60N 30N 0 21S Latitude (°) 34-74S 73S B 10 4 JHIOx+H2SO4+NH3 JH2SO4+NH3 103 102 10 1 10 0 H2SO4 (cm-3) 108 8 6 4 2 107 8 6 4 2 106 0.1 6 8 2 Värriö: April-Oct Helsinki: July-Aug Réunion: April Southern Ocean: Feb-Mar Aboa: Dec-Jan -10 °C 10 °C HIOx+H2SO4 HIOx+H2SO4+NH3 4 6 8 10 2 4 6 8 2 1 HIO3 : H2SO4 ratio Fig. 4. HIOx enhancement of H2SO4-NH3 nucleation. (A) Box plot statistics of HIO3:H2SO4 ratios measured around the globe, showing median values with 10 and 90% percentiles in the whiskers. The gray labels show the site name and months, with data covering >50% of the days. (B) Nucleation enhancement by HIOx is calculated by dividing the measured JHIOx(cid:1)H2SO4(cid:1)NH3 to predict JH2SO4(cid:1)NH3 using CLOUD parameterizations (14). The median ratios at all sites are greater than 0.1, which infers at least a 10-fold nucleation rate enhancement by HIOx. The enhancement is especially pronounced in polar regions where the HIO3:H2SO4 ratio is consistently higher than 0.1. Thin symmetric error bars represent one standard deviation during the data selection periods. In the experiments without NH3 injection (hollow markers), the NH3 concentrations were below the instrument detection limit (4 pptv), which is adopted as a conservative estimate of JH2SO4(cid:1)NH3. However, the actual NH3 concentration is expected to be below 1 pptv, as all charged clusters are essentially NH3 free (Fig. 3). The thick asymmetric error bars represent the systematic uncertainty assuming NH3 equals 1 pptv. The NH3 concentrations in experiments with NH3 injection (filled markers) are well measured, and thus without asymmetric errors. The field observation sites are summarized in the methods. of other compounds, such as MSA and oxi- dized organic molecules, can also contribute to early particle growth, in addition to H2SO4 and HIO3. Climate implications Atmospheric observations show that both iodine oxoacid and sulfuric acid–ammonia nucleation can be important particle sources in specific regions of the pristine boundary layer (6, 17, 18, 20, 21). So far, HIO3 and HIO2 have been thought to be important only in regions where they are more abundant than H2SO4. In polar and marine environments, it is currently thought that H2SO4-NH3 constitutes the primary source of new particle formation, despite the perceived scarcity of NH3 (15). This picture is challenged by our findings. Our data support the reverse: H2SO4-NH3 nucleation plays a major role only when H2SO4 is sub- stantially more abundant than HIO3 and HIO2. The role of HIOx in atmospheric aerosol nuclea- tion may have been overlooked, as studies could easily be deceived by relatively higher H2SO4 than HIOx in parts of the pristine atmosphere. To assess the atmospheric importance of HIOx-H2SO4(-NH3) nucleation, we calculated the J1.7 enhancement factor [the ratio of J1.7 from HIOx-H2SO4(-NH3) to that from H2SO4- NH3] (14) as a function of the HIO3:H2SO4 concentration ratio (Fig. 4B). The enhance- ment factors are large, ranging from 10 to 104 for atmospherically relevant HIO3:H2SO4 ra- tios. Even when the HIO3:H2SO4 ratio is 0.1, the enhancement factor is 10. Observations at marine and polar sites from the North Pole to Antarctica show median HIO3:H2SO4 ratios larger than 0.1 (Fig. 4A), implying that syn- ergistic HIOx-H2SO4(-NH3) nucleation may have global importance and yet has hitherto been overlooked. This conclusion is supported by our calculations of sulfuric acid nucleation enhanced by HIOx, which are shown in fig. S10. At −10°C, which is representative of the marine free troposphere, fast nucleation rates of up to 10 cm−3 s−1 are estimated for ambient acid concentrations. The pronounced temper- ature dependence of HIOx-H2SO4(-NH3) nu- cleation that we find in our study may help explain why nucleation in the marine bound- ary is rarely observed, whereas nucleation is frequently found in the free troposphere or the upper marine boundary layer after passage of a cold front (23, 34, 35). New particle formation from HIOx-H2SO4 has notable implications for the future cli- mate. Iodine oxoacids may enhance CCN and cloud formation in the Arctic (20), which would, in turn, affect both long- and shortwave radia- tive forcing at the surface (36). The absence of iodine oxoacid nucleation mechanisms in climate models may help explain why they systematically underestimate the CCN number concentration around the coast of Antarctica (37, 38). Iodine has also been observed in both gas and particle phases in the polar and ma- rine free troposphere and the upper tropo- sphere and lower stratosphere (39, 40). These regions are characterized by low temperatures and extremely low NH3 concentrations (15), conditions that strongly favor HIOx-H2SO4 or pure HIOx nucleation over H2SO4-NH3 nu- cleation. While global anthropogenic SO2 emissions continue to fall as a result of emission policies, iodine emissions have tripled since the 1950s, and this trend continues (41, 42). As a result, nucleation mechanisms involv- ing iodine oxoacids are anticipated to become even more important in future. To sharpen the understanding of marine aerosol-cloud radia- tive forcing, it is important that representa- tions of new particle formation in global climate models now include iodine oxoacids together with sulfuric acid. REFERENCES AND NOTES 1. V. Masson-Delmotte et al., Eds., Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge Univ. Press, 2021). 2. V.-M. Kerminen et al., Environ. Res. Lett. 13, 103003 (2018). J. Almeida et al., Nature 502, 359–363 (2013). 3. 4. M. Xiao et al., Atmos. Chem. Phys. 21, 14275–14291 (2021). J. Kirkby et al., Nature 476, 429–433 (2011). 5. He et al., Science 382, 1308–1314 (2023) 15 December 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E Corrected 5 January 2024. See full text. 6. T. Jokinen et al., Sci. Adv. 4, eaat9744 (2018). 7. 8. S. M. Kreidenweis, J. H. Seinfeld, Atmos. Environ. 22, 283–296 J. Brean et al., Nat. Geosci. 14, 383–388 (2021). (1988). 9. P. H. McMurry, J. Colloid Interface Sci. 78, 513–527 (1980). 10. C. Kuang, P. H. McMurry, A. V. McCormick, F. L. Eisele, 34. K. J. Sanchez et al., Atmos. Chem. Phys. 21, 3427–3446 (2021). 35. G. Zheng et al., Nat. Commun. 12, 527 (2021). 36. T. Mauritsen et al., Atmos. Chem. Phys. 11, 165–173 (2011). 37. I. L. McCoy et al., Proc. Natl. Acad. Sci. U.S.A. 117, 18998–19006 (2020). J. Geophys. Res. 113, D10209 (2008). 38. L. A. Regayre et al., Atmos. Chem. Phys. 20, 10063–10072 11. C. Yan et al., Geophys. Res. Lett. 48, e2020GL091944 (2020). (2021). 12. R. Wood, Mon. Weather Rev. 140, 2373–2423 (2012). 13. M. L. Dawson et al., Proc. Natl. Acad. Sci. U.S.A. 109, 18719–18724 (2012). 14. E. M. Dunne et al., Science 354, 1119–1124 (2016). 15. B. A. Nault et al., Commun. Earth Environ. 2, 93 (2021). 16. A. Ranjithkumar et al., Atmos. Chem. Phys. 21, 4979–5014 (2021). 17. M. Sipilä et al., Nature 537, 532–534 (2016). 18. X.-C. He et al., Science 371, 589–595 (2021). 19. X.-C. He et al., Aerosol Sci. Technol. 55, 231–242 (2021). 20. A. Baccarini et al., Nat. Commun. 11, 4924 (2020). 21. L. J. Beck et al., Geophys. Res. Lett. 48, e2020GL091334 (2021). 22. H. Finkenzeller et al., Nat. Chem. 15, 129–135 (2023). 23. A. Baccarini et al., J. Geophys. Res. Atmos. 126, e2021JD035126 (2021). 24. T. Hoffmann, C. D. O’Dowd, J. H. Seinfeld, Geophys. Res. Lett. 28, 1949–1952 (2001). 25. C. D. O’Dowd et al., Nature 417, 632–636 (2002). 26. G. R. Wentworth et al., Atmos. Chem. Phys. 16, 1937–1953 (2016). 27. K. E. Altieri, K. A. M. Spence, S. Smith, Geophys. Res. Lett. 48, e2020GL091963 (2021). 28. R. Zhang et al., Environ. Sci. Technol. 56, 14166–14177 (2022). 29. M. J. McGrath et al., Atmos. Chem. Phys. 12, 2345–2355 (2012). 30. J. C. Gómez Martín et al., Nat. Commun. 11, 4521 (2020). 31. S.-H. Lee et al., Science 301, 1886–1889 (2003). 32. E. R. Lovejoy, J. Curtius, K. D. Froyd, J. Geophys. Res. 109, D0820 (2004). 33. D. Stolzenburg et al., Atmos. Chem. Phys. 20, 7359–7372 (2020). 39. R. Volkamer et al., Atmos. Meas. Tech. 8, 2121–2148 (2015). 40. T. K. Koenig et al., Proc. Natl. Acad. Sci. U.S.A. 117, 1860–1866 (2020). 41. C. A. Cuevas et al., Nat. Commun. 9, 1452 (2018). 42. M. Legrand et al., Proc. Natl. Acad. Sci. U.S.A. 115, 12136–12141 (2018). 43. X.-C. He et al., Iodine oxoacids enhance nucleation of sulfuric acid particles in the atmosphere: data resources, version 1, Zenodo (2023); https://doi.org/10.5281/zenodo.8344386. ACKN OWLED GMEN TS We thank CERN for supporting CLOUD with important technical and financial resources and for providing a particle beam from the CERN Proton Synchrotron. We thank the CSC IT Center for Science in Espoo, Finland, for computing time. Funding: This work was supported by Research Council of Finland ACCC Flagship nos. 337549 and 337552; Research Council of Finland professorship (302958); Research Council of Finland Centre of Excellence nos. 346371, 346372, and 346373; Research Council of Finland project nos. 335844, 334514, 345125, 325656, 316114, 314798, 325647, 353836, 341349, 346372, 296628, and 349659; European Research Council (ERC) project nos. 742206, 714621, 616075, and 101002728; Marie Skłodowska-Curie grant agreement nos. 316662, 701647, 895875, and 764991; National Natural Science Foundation of China grant nos. 42175118 and 22236004; the National Key Research and Development Program of China (2022YFC3701000, Task 1); Swiss National Science Foundation grant nos. 200021_169090, 200021_213071, 206021_198140, 200020_172602, and 20FI20_172622; US National Science Foundation grants nos. AGS-2132089, AGS-1801897, AGS-1801574, AGS-1801280, AGS-2027252, AGS-2215522, AGS-1602086, AGS-1801329, and AGS-2215527; German Federal Ministry of Education and Research (BMBF) project no. 01LK1601A; and Portuguese Science Foundation project CERN/FIS-COM/0028/2019. R.C.T. and M.K. thank the Jane and Aatos Erkko Foundation for funding. M.K. received funding from Prince Albert Foundation contract no. 2859. M.W. thanks the Schmidt Science Fellowship for support. X.-C.H. and M.K. thank the Jenny and Antti Wihuri Foundation for funding. Author contributions: X.-C.H., J.Kir., M.Sip., and M.K. planned the experiments. X.-C.H., M.Sim., B.R., J.S., H.F., D.S., A.B., Y.J.T., M.W., S.A., A.A.P., A.A., R.B., Z.B., L.C., B.C., L.D., J.D., I.E.H., R.C.F., M.G., M.H., V.H., W.K., J.Kre., A.K., H.L., B.L., N.G.A.M., V.M., H.E.M., G.M., R.M., D.M., R.L.M., B.M., A.O., T.P., J.P., M.P., M.P.R., S.S., W.S., B.S., M.Sur., A.T., A.C.W., D.W., Y.W., S.K.W., A.W., P.M.W., M.Z.-W., J.Kir., U.B., K.L., J.C., R.V., M.Sip., and M.K. prepared the CLOUD facility or measuring instruments. X.-C.H., M.Sim., S.I., B.R., J.S., H.F., D.S., A.B., Y.J.T., A.A., R.B., L.C., L.D., J.D., I.E.H., M.G., M.H., V.H., T.J., D.K., W.K., J.Kre., H.L., B.L., N.G.A.M., V.M., H.E.M., G.M., R.M., D.M., R.LM., B.M., J.P., A.R., W.S., R.C.T., A.C.W., D.W., S.K.W., M.Z.-W., J.Kir., J.C., and R.V. collected the data. X.-C.H., S.I., H.-B.X., H.F., D.S., R.Z., Y.J.T., L.C., L.D., R.C.F., F.M., W.S., R.C.T., S.K.W., P.M.W., T.K. and D.R.W. analyzed the data. X.-C.H., N.M.D. and J.Kir. wrote the manuscript with contributions from S.I., H.-B.X., B.R. and R.Z. X.-C.H., M.Sim., S.I., H.-B.X., B.R., J.S., H.F., D.S., R.Z., A.B., L.D., J.D., I.E.H., R.C.F., A.H., H.L., N.G.A.M., T.P., S.S., W.S., R.C.T., A.C.W., J.Kir., U.B., K.L., J.C., T.K., D.R.W., R.V., N.M.D., M.Sip. and M.K. commented on and edited the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: Data for all figures in the main text and supplementary materials are available at the Zenodo repository (43). Correspondence and additional requests for materials should be addressed to X.-C.H. (institutional email: xucheng.he@helsinki.fi; permanent email: xuchenghe93@gmail.com). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh2526 Materials and Methods Figs. S1 to S10 References (44–76) Submitted 20 February 2023; accepted 13 November 2023 10.1126/science.adh2526 He et al., Science 382, 1308–1314 (2023) 15 December 2023 6 of 6
10.1126_science.adh4419
RES EARCH R E S E A R C H A R T I C L E ◥ INORGANIC CHEMISTRY Diberyllocene, a stable compound of Be(I) with a Be–Be bond Josef T. Boronski*, Agamemnon E. Crumpton, Lewis L. Wales, Simon Aldridge* The complex diberyllocene, CpBeBeCp (Cp, cyclopentadienyl anion), has been the subject of numerous chemical investigations over the past five decades yet has eluded experimental characterization. We report the preparation and isolation of the compound by the reduction of beryllocene (BeCp2) with a dimeric magnesium(I) complex and determination of its structure in the solid state by means of x-ray crystallography. Diberyllocene acts as a reductant in reactions that form beryllium-aluminum and beryllium-zinc bonds. Quantum chemical calculations indicate parallels between the electronic structure of diberyllocene and the simple homodiatomic species diberyllium (Be2). B ecause of the extreme biotoxicity of be- ryllium, its chemistry is the least well- developed of all the nonradioactive elements (1–3). This toxicity is a mani- festation of beryllium’s particular chem- ical properties: It forms the smallest of all metal ions (Be2+; ionic radius, 0.31 Å; com- pare with that of Li+, 0.60 Å), with unparal- leled charge density (6.45 Å−1; compare with that of Li+, 1.67 Å−1). Its bonding interactions often feature considerable covalent character on account of the highly polarizing properties of the dication (2). Given this capacity for co- valency, the fundamental nature of beryllium- beryllium bonding has been debated for more than a century (4–11). Basic molecular orbital theory predicts that diberyllium (Be2) should have a bond order of zero, and only recently has meaningful insight into the true nature of the beryllium-beryllium interaction in gaseous Be2 been obtained (12, 13). Moreover, despite seminal reports of zinc-zinc and magnesium- magnesium bonds in the early 2000s and a host of quantum chemical investigations into the potential stability of beryllium-containing analogs, the isolation of a compound that fea- tures a beryllium-beryllium bond in the con- densed phases has not been achieved (14–17). Given the dearth of compounds that feature beryllium-beryllium bonds, in recent years, synthetic studies have focused on the prep- aration of low-valent beryllium species sta- bilized by redox noninnocent carbene ligands (18–20). Compounds proposed to contain beryl- lium in the 0 and +1 oxidation states have been synthesized. However, such descriptions have proven controversial and are a source of on- going debate within the literature (21–24). We recently reported the beryllium-aluminyl com- Chemistry Research Laboratory, Department of Chemistry, Oxford, OX1 3TA, UK. *Corresponding author. Email: josef.boronski@sjc.ox.ac.uk (J.T.B.); simon.aldridge@chem.ox.ac.uk (S.A.) plex CpBeAl(NON) [1; NON, 4,5-bis(2,6-diisopro- pylanilido)-2,7-di-tert-butyl-9,9-dimethylxanthene; Cp, cyclopentadienyl anion, [C5H5]−], which displays the reactivity expected from low- valent, nucleophilic beryllium (25–27). In this work, we present diberyllocene, CpBeBeCp (2), an unambiguous beryllium(I) compound that is stable in the condensed phases and features a Be–Be bond. The experimentally determined properties of diberyllocene (which is stable at 80°C for extended periods) validate past quantum chemical predictions of its stab- ility with respect to disproportionation to Be(0) and BeCp2 (7). Synthesis and characterization We previously succeeded in using beryllocene BeCp2 for the synthesis of beryllium-aluminyl complex 1 (25, 28). Calculations indicate that compound 1 features a higher partial positive charge at aluminum than beryllium. Thus, the formation of 1 could be viewed as a reduction of beryllium(II) to beryllium(I) by the aluminyl anion, with associated one-electron oxidation of aluminum(I), which forms a covalent Be–Al bond. We therefore set out to examine the possibility of reducing beryllocene to form a covalent Be–Be bond between two beryl- lium(I) centers. We turned our attention to di-magnesium(I) reagent [(MesNacnac)Mg]2 {3;MesNacnac, [(MesNCMe)2CH]−; Mes, 2,4,6- trimethylphenyl}, which has been used for the controlled reduction of various organo- metallic compounds (6, 17). At room temper- ature, under an inert atmosphere, the reaction of one equivalent of 3 with two equivalents of BeCp2 in toluene leads to quantitative for- mation of diberyllocene CpBeBeCp (2) and (MesNacnac)MgCp, as evidenced by multi- nuclear nuclear magnetic resonance (NMR) spectroscopy (Fig. 1 and fig. S5). Both com- pounds are highly soluble in alkane solvents, such as pentane and hexane. However, 2 is a volatile solid and can be purified through sub- limation at room temperature, which allows for its isolation in 85% yield. Compound 2 can be crystallized from a con- centrated hexane solution, which yields crys- tals suitable for the elucidation of its structure by means of single-crystal x-ray diffraction (Fig. 1B). Compound 2 is shown to feature two half-sandwich (cyclopentadienyl)beryllium units linked through a beryllium–beryllium bond. Zinc is the only other element for which the dimetallocene structural motif has been reported—for example, in Cp*ZnZnCp* [4; Cp*, pentamethylcyclopentadienyl anion (C5Me5)−]— and the configuration of 2 closely resembles that of 4 (14, 15). Both 2 and 4 display D5h symmetry in the solid state, with the cyclo- pentadienyl ligands adopting a parallel, eclipsed configuration. At 2.0545(18) Å, where the num- ber in parentheses is the standard deviation of the least significant digits, the Be1–Be2 distance in 2 is in line with the sum of the single-bond covalent radii for beryllium (2.04 Å) (29). This value is extremely close to that predicted by previous computational investigations of this compound (2.041 to 2.077 Å) (7–9). Examples of hydride-bridged beryllium compounds of the form XBe(m-H)2BeX all feature wider Be–Be separations [range, 2.098(3) to 2.212(8) Å; mean, 2.136 Å] (26, 30–32). The Be–C distances within 2 range from 1.923(2) to 1.938(2) Å (mean, 1.930 Å) and are considerably longer than that of BeCp2 [range, 1.862(9) to 1.936(8) Å; mean, 1.904 Å] (28). Consistently, the Be-(h5-C5H5) centroid distances for Be1 and Be2 are both 1.519 Å, which is again longer than that re- ported for BeCp2 (1.485 Å). These bond metrics reflect the lower oxidation state and greater ionic radius of the beryllium(I) centers in 2 compared with those of the beryllium(II) center in BeCp2. Compound 2 was investigated by use of multinuclear NMR spectroscopy. The 1H NMR spectrum of 2 consists of a single resonance at 5.73 parts per million (ppm), which corre- sponds to the five equivalent cyclopentadienyl ligand protons. The 1H NMR resonances of terminal beryllium-bound hydride ligands are generally found between 4 and 5 ppm (32, 33). Thus, 2 shows no signals attributable to Be–H hydrides in its 1H NMR spectrum, only the resonance that corresponds to the cyclopenta- dienyl ligand. Similarly, the 13C NMR spec- trum of 2 features one signal at 102.7 ppm. Compound 2 was also investigated by means of 9Be NMR spectroscopy—which offers a con- venient probe of the electron density at the beryllium center—and shows a single high-field resonance at −27.6 ppm (25, 34). The coordi- nation of strongly electron-donating anionic ligands to beryllium generally shifts the 9Be NMR resonances further upfield as the in- creasingly electron-rich beryllium center be- comes more strongly shielded. So, for example, CpBeCl exhibits a d9Be shift of −19.5 ppm, Boronski et al., Science 380, 1147–1149 (2023) 16 June 2023 1 of 3 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Synthesis, crystallographic, and electronic structure of diberyllocene (2). (A) Synthesis of 2 through reaction of magnesium(I) complex 3 and beryllocene. (B) Molecular structure of 2 in the solid state, as determined with x-ray crystallography. Thermal ellipsoids set at 50% probability. (C) Calculated HOMO for 2 [isovalue, 0.03 arbitrary units (a.u.)]. whereas the corresponding signals of CpBeMe, CpBeGa(NON), CpBe[Si(CH3)3], and beryllium– aluminyl compound 1 are found at −20.5, −26.9, −27.7, and −28.8 ppm, respectively (table S2) (34). These data suggest that the beryllium center in 2 is similarly as electron rich as that in 1, which is consistent with the description of 2 as a metal-metal bonded beryllium(I) com- pound (25, 34). Moreover, the 9Be NMR shift for compound 2 also provides evidence against a bridging hydride formulation for this com- pound because such a species would be ex- pected to show a much lower field chemical shift (fig. S23) (34). To further argue against the hydride for- mulation for 2, we measured attenuated total reflection infrared (ATR IR) spectra for sam- ples of 2 prepared in both protio and perdeu- tero toluene solvents (figs. S14 and S15). The spectra are identical, and neither features an absorbance band in the 1500 to 2000 cm−1 region where a Be–H stretch might be ex- pected to be observed (35). The infrared spec- tra of both 2 and the hypothetical complex CpBe(m-H)2BeCp were simulated through quan- tum chemical calculations (figs. S17 and S18). The experimentally derived ATR IR spectrum of 2 closely matches the spectrum calculated for this complex (fig. S19) and differs substan- tially from that of the hydride-bridged species (fig. S20), which features a very intense band at 1527 cm−1 [us(Be–Hb)]. The beryllium(II) hydride compound (CpBeH)n is thermally frag- ile and has been reported to undergo ligand redistribution to form BeCp2 and BeH2 within minutes at −10°C (35, 36). This reactivity con- trasts with the relative stability of 2, which shows no obvious signs of degradation upon heating in solution at 80°C for 48 hours. Quantum chemical investigations Quantum chemical calculations were performed on 2 (B3LYP D3BJ def2-TZVP def2/J), in light of the experimental data that had been un- available for previous theoretical investigations of this molecule (7–9). Key structural param- eters align well with those determined crys- tallographically [for example, d(Be–Be) = 2.046 Å, as compared with 2.0545(18) Å]. These calculations indicate that the energetic separation between the highest occupied mo- lecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) is 5.541 eV. The HOMO of 2 is calculated to be the Be–Be s-bonding orbital, with some Be–Cp s-bonding character (Fig. 1C); the lower-lying HOMO − 1 to HOMO − 4 correspond to the Cp–Be p-bonding combinations. As with the homodiatomic mo- lecule Be2, we calculated the Be–Be s* anti- bonding orbital (LUMO + 1) of diberyllocene to be markedly lower in energy than the Be–Be p-bonding orbitals (LUMO + 4 and LUMO + 5) (figs. S27 and S29) (12, 13). This implies that further reduction of 2 would lead to a weakening of the Be–Be interaction. Indeed, the Be–Be distance in Be2 (2.45 Å) is considerably longer than that measured for 2 [2.0545(18) Å] (12, 13). As such, the core of 2 could be considered to be a (cyclopentadienyl- stabilized) dication of diberyllium, [BeBe]2+, which molecular orbital theory would predict to feature a beryllium–beryllium s-bond and a formal bond order of one, as is found here (37, 38). We used natural bond orbital (NBO) calcula- tions to examine the bonding and charge dis- tribution within 2. Natural population analysis (NPA) indicates that the 2s valence orbital of each beryllium atom is substantially populated [0.98 electrons (e–)], with a non-negligible pop- ulation of the beryllium 2p orbitals (0.17 e–). These data are consistent with a formal beryl- lium(I) oxidation state. NBO calculations sug- gest that the Be–Be bond comprises 93% 2s character and 7% 2p character, with each be- ryllium center contributing equally (50:50). We calculated a Wiberg bond index (WBI) of 0.90 for the Be–Be bonding interaction, which is slightly greater than the WBI calculated for the Al–Be bond of 1 (0.82) (25). The charge distribution in 2 was probed with both NPA and quantum theory of atoms in molecules (QTAIM) calculations. QTAIM analysis indi- cates that a non-nuclear attractor is present for the Be–Be interaction, as has been indi- cated with previous theoretical studies of 2, and yields a charge of +1.33 for each beryllium atom (8). The NPA charges at each beryllium center are lower (+0.84), but these do not take into account the presence of the non-nuclear attractor. QTAIM calculations previously per- formed on 1 provide a charge of +1.39 for the beryllium center in this compound, which sug- gests that the beryllium centers in 1 and 2 are similarly electron rich. This is consistent with the similar 9Be NMR shifts measured for the two compounds (25, 34). Because of the high quality of the x-ray crystallographic data obtained for 2, we used quantum-crystallographic methods to gain fur- ther evidence that this compound does not feature bridging hydride ligands. Even with- out using quantum-crystallographic methods, there is insufficient residual electron density to account for the presence of bridging hy- dride ligands (figs. S34 and S35). We employed Boronski et al., Science 380, 1147–1149 (2023) 16 June 2023 2 of 3 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Reactivity of compound 2 with metal-iodide complexes, which leads to the formation of beryllium-metal bonded complexes 1 and 5. iPr, isopropyl; tBu, tert-butyl. the NoSpherA2 method; this technique allows for refinement of crystallographic data on the basis of the partitioned wave function of a molecule, rather than the traditional approach of refining against a model in which electron density is considered as point charges asso- ciated with specific atoms (independent elec- tron model) (39, 40). By using NoSpherA2, essentially all electron density in the data col- lected for 2 is accounted for (figs. S31 and S32). The electron density associated with the cyclo- pentadienyl C–H bonds can be clearly identi- fied (fig. S35), and there is no residual electron density resulting from bridging hydride ligands (figs. S31 and S33). Reactivity studies In view of the metal–metal bond in 2, we probed its reactivity as a source of low-valent beryl- lium. Complex 2 does not react with H2 (one atmosphere), even at elevated temperatures— a finding that is consistent with di-magnesium (I) compounds such as 3, which also do not react with H2 (17). However, compound 2 re- duces (NON)AlI to cleanly and quantitatively yield the known beryllium-aluminyl compound 1 and CpBeI (Fig. 2, left), which evidences the usefulness of 2 for the synthesis of beryllium- metal bonds (25). This reactivity would not be expected from a beryllium(II) hydride com- pound. The reaction of 2 with (DippNacnac)ZnI (DippNacnac, [(DippNCMe)2CH]−; Dipp, 2,6- diisopropylphenyl) was similarly tested to ex- amine whether the formation of a (hitherto unknown) Zn–Be bond could be achieved (Fig. 2, right). In a similar fashion, 1H NMR spec- troscopy indicates the complete consumption of 2, along with quantitative formation of CpBeI and a (DippNacnac)-containing species. Additionally, the formation of a small amount of dark gray metallic precipitate was observed. Crystallization from hexane yielded CpBeZn (DippNacnac) (5), which was structurally au- thenticated by single-crystal x-ray diffraction (fig. S21). The Zn–Be bond within compound 5 repre- sents a very rare example of a beryllium-metal bonding combination and provides further evi- dence for the constitution of 2 itself (25, 27, 41). The Zn–Be bond distance within 5 [2.169(10) Å] is in line with the sum of the covalent radii of Be and Zn (2.20 Å) and is consistent with the presence of a covalent metal-metal bond (29). Additionally, 5 exhibits a 9Be NMR resonance (−27.7 ppm) that is consistent with a very electron-rich beryllium metal center, similar to those of 1 and 2 (25, 34). Quantum chemical calculations performed on 5 (B3LYP D3BJ def2-TZVP def2/J) yield a similar Zn–Be dis- tance (2.175 Å) to that determined crystallo- graphically and indicate that the HOMO is a Zn–Be s-bonding orbital (fig. S30). On the basis of the Pauling electronegativities of be- ryllium (1.57) and zinc (1.65), compound 5 might be assigned a beryllium(II)/zinc(0) formalism. However, quantum chemical calculations indi- cate considerable covalent contributions to the Zn–Be bonding in this compound, in simi- lar fashion to the Be–Al bonding in 1 (25). For example, NPA calculations imply that the va- lence orbitals of both zinc and beryllium are substantially populated (Be, 2s, 1.00 e– and 2p, 0.16; Zn, 4s, 1.04 e–). Additionally, NBO analy- sis suggests that beryllium and zinc make almost equal contributions to the Zn–Be bond (Zn:Be, 49.9:50.1). Thus, a beryllium(I)–zinc(I) formulation seems a more appropriate descrip- tor for 5. After half a century, diberyllocene (2) has been synthesized. The Be–Be distance in 2 [2.0545(18) Å] is in line with all previous quan- tum chemical investigations of this com- pound. Moreover, 2 reacts as a reductant and can be used to synthesize beryllium-metal bonds. RE FERENCES AND NOTES 1. R. Puchta, Nat. Chem. 3, 416 (2011). 2. D. Naglav, M. R. Buchner, G. Bendt, F. Kraus, S. Schulz, Angew. Chem. Int. Ed. 55, 10562–10576 (2016). 3. M. R. Buchner, Chem. Commun. 56, 8895–8907 (2020). 4. C. Jones, Commun. Chem. 3, 159 (2020). 5. L. A. Freeman, J. E. Walley, R. J. Gilliard Jr., Nat. Synth. 1, 439–448 (2022). 6. S. J. Bonyhady et al., Chemistry 16, 938–955 (2010). 7. A. Velazquez, I. Fernández, G. Frenking, G. Merino, Organometallics 26, 4731–4736 (2007). 8. X. Li et al., Organometallics 32, 1060–1066 (2013). 9. Y. Xie, H. F. SchaeferIII, E. D. Jemmis, Chem. Phys. Lett. 402, 414–421 (2005). 10. S. A. Couchman, N. Holzmann, G. Frenking, D. J. D. Wilson, J. L. Dutton, Dalton Trans. 42, 11375–11384 (2013). 11. H. Schmidbaur, Be Organoberyllium Compounds (vol. 21, Gmelin Handbook of Inorganic and Organometallic Chemistry, Springer, 1987). 12. J. M. Merritt, V. E. Bondybey, M. C. Heaven, Science 324, 1548–1551 (2009). 13. K. Patkowski, V. Špirko, K. Szalewicz, Science 326, 1382–1384 (2009). I. Resa, E. Carmona, E. Gutierrez-Puebla, A. Monge, Science 14. 305, 1136–1138 (2004). 15. A. Grirrane et al., J. Am. Chem. Soc. 129, 693–703 (2007). 16. S. P. Green, C. Jones, A. Stasch, Science 318, 1754–1757 (2007). 17. C. Jones, Nat. Rev. Chem. 1, 0059 (2017). 18. M. Arrowsmith et al., Nat. Chem. 8, 638–642 (2016). 19. C. Czernetzki et al., Angew. Chem. Int. Ed. 60, 20776–20780 (2021). 20. G. Wang et al., J. Am. Chem. Soc. 142, 4560–4564 (2020). 21. D. Jędrzkiewicz et al., Angew. Chem. Int. Ed. 61, e202200511 (2022). 22. M. Gimferrer et al., Chem. Sci. 13, 6583–6591 (2022). 23. M. Gimferrer et al., Chem. Sci. 14, 384–392 (2023). 24. S. Pan, G. Frenking, Chem. Sci. 14, 379–383 (2022). 25. J. T. Boronski et al., J. Am. Chem. Soc. 145, 4408–4413 (2023). 26. A. Paparo et al., Dalton Trans. 50, 7604–7609 (2021). 27. A. Paparo, C. D. Smith, C. Jones, Angew. Chem. Int. Ed. 58, 11459–11463 (2019). 28. R. Fernández, E. Carmona, Eur. J. Inorg. Chem. 2005, 3197–3206 (2005). 29. P. Pyykkö, J. Phys. Chem. A 119, 2326–2337 (2015). 30. M. Arrowsmith, M. S. Hill, G. Kociok-Köhn, Organometallics 34, 653–662 (2015). 31. G. W. Adamson, H. M. M. Shearer, Chem. Commun. 11, 240 (1965). 32. T. J. Hadlington, T. Szilvási, Nat. Commun. 13, 461 (2022). 33. R. Han, G. Parkin, Inorg. Chem. 31, 983–988 (1992). 34. P. G. Plieger et al., J. Am. Chem. Soc. 126, 14651–14658 (2004). 35. M. del Mar Conejo et al., Chemistry 9, 4462–4471 (2003). 36. T. C. Bartke, A. Bjorseth, A. Haaland, K.-M. Marstokk, H. Møllendal, J. Organomet. Chem. 85, 271–277 (1975). 37. C. A. Nicolaides, P. Valtazanos, Chem. Phys. Lett. 174, 489–493 (1990). 38. A. Kalemos, J. Chem. Phys. 145, 214302 (2016). 39. L. Midgley et al., Acta Crystallogr. A Found. Adv. 77, 519–533 (2021). 40. F. Kleemiss et al., Chem. Sci. 12, 1675–1692 (2020). 41. H. Braunschweig, K. Gruss, K. Radacki, Angew. Chem. Int. Ed. 48, 4239–4241 (2009). 42. J. T. Boronski, A. E. Crumpton, L. L. Wales, S. Aldridge, Diberyllocene: xyz-coordinates of optimized structures. Dryad (2023); https://doi.org/10.5061/dryad.kwh70rz8n. AC KNOWLED GME NTS Funding: J.T.B. thanks St John’s College, Oxford, for a Junior Research Fellowship. We thank the John Fell Fund (0011792) for financial support. We thank the EPSRC Centre for Doctoral Training in Inorganic Chemistry for Future Manufacturing (OxICFM, EP/ S023828/1 studentship for L.L.W. and A.E.C.). Author contributions: J.T.B.: conceptualization, investigation, visualization, writing – original draft, writing – review and editing, funding acquisition, supervision, and project administration. A.E.C.: quantum-crystallographic investigations, writing – review and editing. L.L.W.: collection and processing of x-ray crystallographic data, writing – review and editing. S.A.: supervision, project administration, and writing – review and editing. Competing interests: The authors declare no competing interests. Data and materials availability: X-ray data are available free of charge from the Cambridge Crystallographic Data Centre under reference numbers CCDC 2245595 (2), 2245596 (5), and 2246346 [(MesNacnac)MgCp]. All other experimental, spectroscopic, crystallographic, and computational data are included in the supplementary materials. Computational data are also available through Dryad (42). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh4419 Materials and Methods Figs. S1 to S35 Tables S1 to S3 References (43–56) Submitted 6 March 2023; accepted 25 April 2023 10.1126/science.adh4419 Boronski et al., Science 380, 1147–1149 (2023) 16 June 2023 3 of 3
10.1126_science.adg6881
RES EARCH METALENSES Extreme ultraviolet metalens by vacuum guiding Marcus Ossiander1*†, Maryna Leonidivna Meretska1†, Hana Kristin Hampel2†, Soon Wei Daniel Lim1, Nico Knefz2, Thomas Jauk2, Federico Capasso1*, Martin Schultze2* Extreme ultraviolet (EUV) radiation is a key technology for material science, attosecond metrology, and lithography. Here, we experimentally demonstrate metasurfaces as a superior way to focus EUV light. These devices exploit the fact that holes in a silicon membrane have a considerably larger refractive index than the surrounding material and efficiently vacuum-guide light with a wavelength of ~50 nanometers. This allows the transmission phase at the nanoscale to be controlled by the hole diameter. We fabricated an EUV metalens with a 10-millimeter focal length that supports numerical apertures of up to 0.05 and used it to focus ultrashort EUV light bursts generated by high-harmonic generation down to a 0.7-micrometer waist. Our approach introduces the vast light-shaping possibilities provided by dielectric metasurfaces to a spectral regime that lacks materials for transmissive optics. D ielectric metasurfaces consist of trans- parent nanostructures with subwave- length separation, which manipulate the phase of light on the nanoscale (1). This elaborate control is revolutioniz- ing modern optics: Metasurfaces can replace bulk optics by thin and flat elements (2, 3), combine multiple functions in single optical elements (4, 5), and be used to realize inno- vative optical components that induce, for example, freely designable optical angular momentum (6) and polarization (7, 8). Tech- nology, including modern semiconductor lithography, demands this design liberty for innovative optical elements for ever-shorter wavelength radiation, but this development has been stalled at ultraviolet frequencies where dielectrics stop being transparent. To our knowledge, linear metaoptics have only been demonstrated down to a wavelength of ≈250 nm (9–11). Nonlinear metasurfaces reach further into the ultraviolet spectrum at the cost of indirect light-shaping mechanisms and have, at present, been demonstrated down to a wave- length of 185 nm (12–15). Inaccessible to metasurface design has been extreme ultraviolet radiation (EUV), which covers the wavelength range from 10 to 121 nm and corresponds to a photon energy of 10 to 124 eV (16). This wavelength regime receives appreciable attention as a gateway to achiev- ing attosecond temporal resolution in ultrafast spectroscopy (17) and lithographically fabricat- ing nanometer-scale transistors in state-of-the- art semiconductor industry (18). However, today, all fields that use EUV radiation are encumbered by handling problems that arise from being limited to reflective or binary optics [e.g., to- 1John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA. 2Institute of Experimental Physics, Graz University of Technology, 8010 Graz, Austria. *Corresponding author. Email: mossiander@g.harvard.edu (M.O.); capasso@seas.harvard.edu (F.C.); schultze@tugraz.at (M.S.) †These authors contributed equally to this work. roidal mirrors or Fresnel zone plates; see (19) for an overview of existing technology]. Here, we present a new physical mechanism for meta- surface design and demonstrate how linear metasurfaces can be realized at a wavelength of 50 nm, thus providing the foundation for general-purpose transmissive optics technology for EUV radiation. Principle of vacuum guiding and metalens design In the EUV, the strong absorption of most materials and their near-unity real part of the refractive indexes (20) usually prevent effective refraction or waveguiding. The refractive in- dexes of most dielectrics in the visible spectrum are determined by electronic transitions in the ultraviolet, that is, by resonances whose frequencies are higher than the frequency w of visible light. In the Drude-Lorentz oscillator model, this results in a complex refractive index ~n wð Þ ¼ n wð Þ þ ik wð Þ for visible light, with a real part n ≥ 1, the imaginary unit i, and a negligi- ble absorption coefficient k. By contrast, EUV light oscillates faster than these electronic resonance frequencies, resulting in n ≤ 1 and a large absorption coefficient k, which renders conventional transmissive metaoptics design unfeasible. For the same reason, EUV manip- ulation must rely on reflective glancing-angle mirrors in vacuum. The concept for metasur- face design introduced here is visualized in Fig. 1: In the EUV spectrum, vacuum or air (n = 1) has a refractive index that is larger than that of a pillar made from n < 1 material; there- fore, the pillar cannot guide or confine light. However, a void or hole (n = 1), that is, the absence of material, in a layer with mate- rial index n < 1 can act as a waveguide core surrounded by a lower-index cladding. There- fore, truncated waveguide metasurfaces are possible in the EUV by following an inverted design scheme that tunes nanohole dimen- sions instead of the shapes of free-standing nanostructures. Materials with n < 1 that can be used to re- alize such metasurfaces exist throughout the EUV; for example, aluminum, silicon, and be- ryllium allow optics for the wavelength range from 40 to 90 nm; scandium and boron cover 20 to 40 nm; and rhenium, molybdenum, and zirconium cover 10 to 20 nm. Figure S1 com- piles the refractive indexes and the trans- mission of these materials. Because of the availability of high-brightness laser-driven tin plasma sources, 13.5 nm is a wavelength of major importance for semiconductor lithog- raphy (18). At this wavelength, for example, ruthenium has the complex refractive index ~n ¼ 0:88 þ 0:02i (21). For the implementation of this concept, we chose a thin membrane of crystalline silicon as the base material and a cylindrical hole as the polarization-independent guiding struc- ture. These are schematically shown (Fig. 1, A and C) together with the real part of the silicon refractive index in the EUV (22) and the trans- mission through a 220-nm-thick silicon layer (Fig. 1B). To highlight the vacuum-guiding behavior of the holes, the simulated intensity profile of light with a vacuum wavelength of lvac = 50 nm incident on such a perforated silicon membrane [80-nm hole diameter in a square 120-nm–by–120-nm unit cell, periodic boundary conditions; see section 1 of (19) for simulation details] is plotted in Fig. 1D: At the center of the membrane (110 nm after its front surface), 84% of the energy of an incident plane wave is transmitted within the hole, whereas 16% of its energy is transmitted in the silicon. However, the hole only covers 34% of the unit- cell area. Because most power is transmitted in vacuum, absorption in silicon is limited, and the overall transmission is enhanced relative to that of the unstructured film: An unstructured 220-nm-thick silicon membrane transmits 28% of incoming 50-nm light. Accounting for the 80-nm-diameter hole using its area coverage would increase transmission to 52%. Vacuum guiding increases the transmission further to 67%. Although perforations have been used in nano-optics before, the presented guiding mechanism is fundamentally different from antiguiding in holes (23), low-index guiding in air (24), or hollow-core fibers (25). Furthermore, the enhancement does not require a periodic structure, which distinguishes the effect from extraordinary optical transmission (26). To realize our EUV metasurface (Fig. 1A), with full design flexibility to emulate the phase profile of a desired optical element, we numer- ically created a library of meta-atoms based on the transmission phase of holes with 20- to 80-nm diameters in a 220-nm-thick silicon membrane [see section 1 of (19) for simulation details]. Notably, between wavelengths of 50 and 62 nm, the photon energy–dependent transmission phase is widely tunable by the Ossiander et al., Science 380, 59–63 (2023) 7 April 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Vacuum-guiding enables EUV metalenses. (A) Concept and simulation of a metalens that is focusing EUV: We impart the phase profile of an aspheric focusing lens on light pulses with a vacuum wavelength of 50 nm (purple disks) using holes through a silicon membrane (rectangular area). Because the refractive index of silicon is smaller than unity in parts of the EUV spectrum, holes through silicon concentrate incoming light. This effect relaxes subwavelength requirements for creating metasurfaces, allows us to impart a hole size– dependent phase shift using feature sizes on the order of the vacuum wavelength, and increases transmission through the absorbing membrane. The false color plot illustrates this light concentration in the holes and how the ultraviolet radiation collapses into a focus after propagating the focal length. For better visibility, we cut the displayed metasurface and the light-intensity distribution along a plane that includes the optical axis. Further simulation details are presented in Fig. 4. (B) Photon energy–dependent real part of the refractive index of crystalline silicon [blue line, data from (22)] in the EUV spectrum and intensity transmission of a 220-nm-thick silicon membrane (red line). The frequencies of the bulk wP and surface plasmon wSP are marked in purple. (C) Schematic and setup for meta-atom simulation: EUV light (purple arrow) passes through a 220-nm-thick crystalline silicon membrane (blue) with a hole with diameter d. We model a single unit cell (120 nm by 120 nm) with periodic boundary conditions. (D) Finite-difference time-domain simulation of EUV vacuum-guiding through an 80-nm-diameter hole in a 220-nm-thick silicon membrane. The false color plot shows the transverse beam intensity profile of light with a vacuum wavelength of 50 nm at the midpoint of the silicon membrane along the propagation direction, that is, 110 nm after the front surface. The hole is indicated as a blue circle. The simulation setup in three dimensions is shown in (C). The hole covers 34% of the total area; however, 84% of the energy is transmitted within the hole and only 16% of the energy is found in silicon. The intensity decays exponentially into the silicon cladding because of the refractive index contrast. The overall transmission of the patterned 220-nm-thick silicon membrane is 67%. Fig. 2. Design and fabrication of an EUV metalens. (A) Photon energy–dependent transmission phase of holes in a 220-nm-thick silicon membrane (see Fig. 1C for the unit cell), color coded for different hole sizes. The gray area indicates the region where hole diameters from 20 to 80 nm offer phase coverage larger than 1.5p, which is enough to achieve efficient and diffraction-limited focusing (27). (B) Overall hole diameter– dependent intensity transmission (purple crosses) and transmission phase in the forward direction (blue circles) of the resulting meta-atom library at 50-nm wavelength (25-eV photon energy). Because the 120-nm–by–120-nm unit-cell size is comparable to the wavelength, low diffraction orders can be generated for holes that cause a transmission phase shift close to p (diameters around 45 nm). When plotting only the transmission into the zeroth diffraction order (red circles) of a periodic array of same-diameter holes, this causes a dip in the transmission. Because hole diameters spatially vary in a metalens and light from all holes interferes constructively to a focus, the more uniform overall transmission (purple crosses) is a better gauge to judge transmission uniformity. (C) Target transmission-phase profile (blue line) of a metalens with focal length f = 10 mm at 50-nm wavelength, calculated from Eq. 1 modulo 2p, and the corresponding matched hole diameter (red circles) using the library of (B) to realize the metalens. (D) Shown at the bottom is a scanning electron microscope (SEM) picture of a 3-mm–by–0.5-mm portion of the metalens with 1-mm diameter and 10-mm focal length designed for 50-nm wavelength. The position where the picture was taken on the metalens is marked by the purple arrow in (G). Shown at the top is the design of the metalens in this area [compare with (C)]. (E) Cross section of a metalens fabricated using the same recipe as the lens in (D) on the silicon-on-insulator carrier wafer obtained using focused-ion-beam milling and SEM. The position where the picture was taken on the metalens is marked by the purple arrow in (G). (F) Zoomed-out SEM picture of a 34-mm–by–31-mm portion of the metalens. The position where the picture was taken on the metalens is marked by the green rectangle in (G). The focusing pattern of the metalens is apparent from the ring segments, which show decreasing width from left to right. Every 10 mm, holes are omitted to increase the stability of the metalens, which is visible as a square scaffolding pattern. The symmetry of the scaffolding is intentionally different from the symmetry of the metasurface to increase stability. (G) Optical microscope picture of the final metalens membrane. The metalens (ML) is encircled by the dashed blue line. Because its features are too small to be resolved at this magnification, the metalens shows a moiré pattern (ring patterns and bright area at the center; an enlarged image is provided in fig. S4). The unpatterned silicon membrane area appears solid gray (encircled by the dashed red line). Areas with a remaining buried oxide layer appear red and green because of thin-film interference (encircled by the dashed yellow line). The silicon carrier wafer appears black. Ossiander et al., Science 380, 59–63 (2023) 7 April 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E hole diameter and offers more than 1.5p trans- mission phase coverage with an average trans- mission of 40% at 50 nm (see Fig. 2A for the photon energy–dependent transmission phase, Fig. 2B for the transmission and transmission phase at 50 nm, and fig. S2 for the photon energy dependent transmission). This transmission- phase coverage is enough to achieve efficient and diffraction-limited focusing, as explored, for example, in (27). The metasurface unit cell is shown in Fig. 1C, and the corresponding library is shown in Fig. 2B. Experimental results To experimentally prove that the vacuum- guiding concept yields viable EUV metalenses, we forward-designed a focusing EUV metasur- face by mimicking the wavelength-dependent transverse hyperbolic phase profile (28) (cid:4) 2p lvac ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi r2 þ f 2 ϕ r; lvac Þ ¼ (cid:2) ð1Þ (cid:2) f p (cid:3) ð p of an aspheric lens with focal length f = 10 mm ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x2 þ y2 in vacuum at transverse position r ¼ (x and y are the cartesian coordinates centered at the beam axis). This analog phase profile is matched by a simulated digital phase profile (sampled at positions x, y; x = kDx; y = lDy; k; l ∈ ℤ; and Dx = Dy = 120 nm; k and l are integer indexes) using the hole library (Fig. 2B), yielding a recipe for the required hole- diameter distribution (Fig. 2C). The metalens is designed for a central vacuum wavelength of lvac = 50 nm, where silicon features a refractive index ~n ¼ 0:77 þ 0:02i. The smaller-than-unity real part at this wavelength partially relaxes the necessity for true subwavelength patterning, which facilitates the manufacturing of the meta- optical element. In the given implementation, the maximum feature size (80 nm) and unit cell size (120 nm) correspond to 1.2 and 1.8 times the inside-silicon wavelength lSi, respectively. Although this does not entirely prevent the for- mation of low diffraction orders that contain up to 53% of the transmitted light, it still allows the realization of numerical apertures up to NA < lvac 2Dx ¼ 0:2 following the Nyquist theorem (29). The demonstration sample, a free-standing metalens with a 1-mm diameter (numerical aperture NAmax = 0.05), is realized from silicon- on-insulator wafers [see section 2 of (19) and fig. S3 for fabrication details]. Figure 2, D and F, shows scanning-electron microscopy pictures of the final sample after metasurface etching but before membrane isolation. Figure 2G shows a light microscopy picture of the finished membrane, with thin-film interference colors confirming the complete removal of the buried oxide layer in the lens area. We achieved the designed hole diameters (see Fig. 2D) using both diameter-dependent electron beam lithog- raphy doses and diameter-dependent fabrication offsets [see section 2 of (19)]. A focused-ion-beam cut (Fig. 2E) through a sample reveals holes Fig. 3. Experimental demonstration of EUV metalens focusing. (A) An intense near-infrared femtosecond laser pulse (red arrow and area) is focused into an argon gas target (green) to generate an attosecond pulse train (purple arrow and area) by high-harmonic generation. Near-infrared radiation is blocked using an aluminum filter foil (gray). The attosecond pulse train is then focused using the metalens (blue) pictured in Fig. 2. At the position of the focus along the propagation direction (marked z), a knife-edge scan is performed using a razor blade mounted on a piezo stage moving along the transverse beam direction (marked x). Afterward, the attosecond pulse train’s spectral components are split using a grazing incidence toroidal grating, and the focal plane is imaged on a charge-coupled device (CCD) camera. (B) EUV beam profile after the metasurface (false color plot) at 25.3-eV photon energy (21st harmonic of the driving laser at 1030-nm wavelength) detected by the CCD. For comparison, we repeat the outlines from the microscopy image in Fig. 2G: The dashed blue line marks the metasurface, and the dashed red line marks the unpatterned silicon area. The granular structure with low intensity is already present in the incoming beam profile, which is plotted in fig. S5. The focal spot created by the metalens, which is imaged onto the CCD using the toroidal grating, is marked by the green dashed rectangle. It appears larger than the real focus because of the limited numerical aperture of the toroidal grating and aberrations caused by the imaging system. (C) Knife-edge scans for different positions along the propagation direction of the metasurface- focused beam [movement direction marked z in (A)]. The colored lines show the knife position–dependent [movement direction marked x in (A)] integrated photon flux detected by the CCD camera in the focus area. As the razor blade moves into the focus, it blocks part of the transmitted radiation and decreases the transmitted flux. A large negative derivative of the flux represents a small focus. The error bars represent the standard deviation of three measurements. The black lines are least-squares fits to the data assuming a Gaussian focus profile. (D) Same as (C) but in close proximity to the focus. The error bars represent the standard deviation of 10 measurements. (E) Propagation direction–dependent waist sizes extracted from the fits in (C) and (D) (blue dots). The error bars represent the 95% confidence interval. The red line is a fit to the waist sizes assuming Gaussian beam propagation. The inset shows a zoomed-in view of the propagation direction–dependent waist size close to the focus [extracted only from the fits in (D)]. The minimum waist sizes reported in the text are marked by the black arrow. Ossiander et al., Science 380, 59–63 (2023) 7 April 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Finite-difference time-domain modeling and benefits of an EUV metalens. (A) Target transverse-phase profile (blue line) of a diffraction-limited metalens (focusing length f = 10 mm, size 6 mm by 6 mm) designed for incoming light with 50-nm wavelength (25-eV photon energy) and sampling of this phase profile with the library presented in Fig. 2 (green crosses). As a comparison, the transmission profile of a binary intensity Fresnel zone plate with the same numerical aperture, focal length, and size is also shown (red line). (B) On the left is a two-dimensional design of a metasurface that realizes the phase profile in (A). White areas represent a 220-nm-thick silicon membrane, and blue areas represent holes through the silicon membrane. On the right is a two-dimensional design of a binary intensity Fresnel zone plate that realizes the transmission profile in (A). White areas are perfectly transmitting, and red areas are perfectly absorbing. (C) Modeled transverse intensity cuts through the focus generated by the metasurface (blue dashed line) and the zone plate in (B) (red line) for incoming light with 50-nm wavelength (25-eV photon energy) and illumination by a Gaussian beam with a 2-mm waist. The zone-plate focus has characteristic side lobes that are not present in the metasurface focus. (D) Shown on the left is the modeled light-intensity evolution (false color plot) after the metasurface pictured in (B) focuses the Gaussian beam described in the caption of (C). Shown on the right is the modeled light- intensity evolution (false color plot) after the zone plate pictured in (B) focuses the same Gaussian beam. with square sidewalls and a partial etch of the smallest diameter holes. Because of the small transmission-phase difference between a mem- brane with small holes and a solid membrane (see Fig. 2B), the resulting phase error is smaller than 0.1p and can be corrected during meta- atom library calculation. For experimental verification of the focusing power of the metalens, we generated diverging EUV attosecond pulse trains through near- infrared femtosecond laser pulse–driven high- harmonic generation in argon gas (30–32) [Fig. 3A and section 3 of (19)]. The frequency up-conversion extends up to the 35th order (42.1-eV photon energy, 29-nm wavelength) of the driving laser pulses (1.2-eV photon energy, 1030-nm wavelength), with spectral power con- centrated around the laser’s odd harmonics. A toroidal EUV grating disperses the spectral components of the attosecond pulse train and creates a frequency-resolved image of the focal plane on an EUV-sensitive camera where the metasurface’s effect at the design wavelength can be inspected. Figure 3B shows the beam profile at the focal plane of the metalens of the 21st harmonic with 25.3-eV photon energy and 49-nm wavelength (close to the design wavelength of the optics). The outline of the circular metasurface (dashed blue line) and features caused by the remaining silica aperture (dashed red and yellow lines; compare with Fig. 2G and fig. S4) are also vis- ible. The bright focal spot at the metasurface center (dashed green line) presents experi- mental evidence for the viability of the EUV metalens to focus incident light. Because the grazing incidence toroidal im- aging grating provides a considerably smaller numerical aperture than the metasurface and introduces aberrations and astigmatism, the obtained image does not determine the focal spot diameter and underestimates the focus- ing power of the optical element. To deter- mine the real focal spot size produced by the metasurface, we implemented a knife-edge scan [see Fig. 3A and (33)], where part of the focused beam in the focal plane is gradual- ly blocked by a razor blade translated along the x direction indicated in Fig. 3A and the position-dependent transmitted intensity is recorded. As focusing concentrates the beam intensity along the transverse direction, the negative spatial derivative of the recorded x-dependent intensity reveals the beam pro- file. Figure 3, C and D, displays scan results for different planes along the propagation di- rection around the focus. Figure 3D includes a knife-edge scan that features a maximum negative spatial derivative, which is indic- ative of the focal plane. Under the assump- tion of a cylindrically symmetric Gaussian beam, the corresponding beam size is ex- tracted by fitting an error function to the data at each position along the propagation direc- tion (Fig. 3E) (33). p ¼ w ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi x2 þ y2 We observe that the metasurface focuses the illuminating beam to a minimum waist of = 0.7 ± 0.3 mm [all reported waists wmetasurface 0 w are measured using the 1/e2 intensity, that is, (cid:6) (cid:5) Þ=e2]. Using I r ¼ the Rayleigh-Sommerfeld diffraction integral (34), we calculated the minimum achieva- = 0.45 mm, assuming ble waist wdiffractionlimit 0 diffraction-limited focusing of our incoming beam (see fig. S5), which highlights that the metalens already performs within 1.6 times of ð ¼ I r ¼ 0 the diffraction limit. For further comparison, the measured propagation distance–dependent waist size w(z) can be fitted to that of a focused Gaussian beam with minimum waist w0 (35) in vacuum: w zð Þ ¼ w0 ð2Þ s (cid:7) ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (cid:8) 2 zlvac pw2 0 1 þ 0 The fit (see Fig. 3E) suggests a minimum pos- sible waist size of wmetasurface = 0.56 ± 0.03 mm, which is even closer to the diffraction limit. Both results overlap within the experimental uncertainty. We attribute the deviation from the diffraction limit to imperfections in the EUV beam guiding and filtering optics and possible residual corrugations of the silicon membrane. For comparison, achieving similar spot sizes using the near-infrared driving laser would require close-to-unity numerical aper- tures; in the EUV, only a numerical aperture of 0.05 is required (36). Aside from the focusing power, the transmis- sion properties are crucial for future applica- tions. Photons with wavelengths shorter than 100 nm possess enough energy to overcome the bandgap of all known dielectrics; there- fore, large absorption is unavoidable (37). None- theless, owing to the vacuum guiding concept, our sample transmits more than 10% of all incoming 49-nm light and focuses 48% of the transmitted 49-nm light, which limits the root- mean-squared metalens wavefront error (38) due to fabrication accuracy to lvac/10 [lvac = 49 nm; see section 4 of (19) for details]. Such fine-granular phase control not only is a pre- requisite for focusing but also opens the door for the future demonstration of optical angular Ossiander et al., Science 380, 59–63 (2023) 7 April 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E momentum plates and general holograms at EUV wavelengths. Simulation of Nyquist-limited focusing To further explore the potential of EUV meta- lenses, we investigated a metalens design with focal length f = 10 mm and overall optics di- ameter d = 6 mm (see Fig. 4A for the phase profile and Fig. 4B for the final design). We then simulated the focusing of a linearly po- larized Gaussian beam using finite-difference time-domain modeling [illuminating Gaussian beam waist willum: = 2 mm, and effective nu- i merical aperture NAeff ¼ sin tan(cid:2)1 willum: willum: f ¼ 0:2 (36), which corresponds to the max- imum realizable numerical aperture given by the Nyquist sampling theorem and our unit cell size (29); see section 1 of (19) for simulation details]. (cid:3) ≈ (cid:4) h f 0 0 Figure 4D shows the formation of the meta- surface focus. Even under these challenging con- ditions, the metasurface focus closely approaches the diffraction limit (wdiffractionlimit = 85 nm) with a minimum beam waist wmetasurface = 94 nm. The metalens focusing properties for a light pulse with extended bandwidth are ex- plored in section 5 of (19) and fig. S6. Having the unit cell size be of the order of the design wave- length causes diffraction of ~53% of the inci- dent power away from the beam axis into the diffraction orders of the quasi-periodic unit-cell arrangement. Adding an unpatterned layer of silicon with a refractive index n = 0.77 and a thickness on the order of half a wavelength after the metalens changes the grating condition in transmission and would prevent the creation of most of these propagating diffraction orders [be- cause it limits the grating indices p; q ∈ ℤ that satisfy the transverse momentum wavevector ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (cid:3) (cid:4) (cid:6)2 þ 2pq (cid:5) 2 2pp Þ2 condition nk0 ¼ Dx with the overall momentum k0 and the mo- mentum along the layer normal kz]. ð þ nkz r Dy , For thorough comparison, we also modeled the focal profile of a binary absorption zone plate with equal numerical aperture (see Fig. 4A for the absorption profile and Fig. 4B for the design). The juxtaposition of the focal profiles generated by the zone plate and the metalens shown in Fig. 4D highlights notable differences in focus quality and corroborates the benefit of the innovative metalens. A comparison with state-of-the-art technology [see (39) for a zone plate with comparable outermost zone width and section 6 of (19) for a summary of EUV focusing optics] highlights that the zone plate creates side lobes in its focal plane, which is an unavoidable property of zone plate foci (40). By contrast, because the metasurface realizes the focusing phase profile accurately by sup- pressing spherical aberrations, no sidelobes are visible. Furthermore, because vacuum guiding decreases absorption and no energy is lost to sidelobes, the maximum intensity in the meta- surface focus exceeds that of the zone plate by 9%. The transverse focal cuts in Fig. 4C high- light this behavior: Unwanted features present in the focal plane are suppressed by more than 10 dB for the metasurface compared with the zone plate. Concluding remarks The transfer of metasurface technology, with its associated superior design freedom to the EUV spectral region, provides a general route to manufacture transmissive optics in this fre- quency range. This capability should lead to applications such as microscopy with unprec- edented spatial and temporal resolution, orbital angular momentum beams with ultrahigh fre- quency, and structured light that has direct access to core-level electronic transitions in atoms and molecules. EUV lithography has become the main enabling fabrication tech- nology that allows us to keep up with Moore’s law (18); conversely, metasurface-based optics can be fabricated with deep ultraviolet lithog- raphy in the same semiconductor foundries of mainstream complementary metal-oxide- semiconductor (CMOS) technology (27). This convergence of semiconductor-processing tech- nology and optics will expand to the realization of metaoptics using EUV lithography, further shrinking feature sizes and increasing the complexity of nanostructure shapes. In turn, with metasurfaces operating in the EUV, they will enable a new generation of lithography optics. RE FERENCES AND NOTES 1. S. M. Kamali, E. Arbabi, A. Arbabi, A. Faraon, Nanophotonics 7, 1041–1068 (2018). 2. N. Yu et al., Science 334, 333–337 (2011). 3. M. Khorasaninejad et al., Science 352, 1190–1194 (2016). 4. N. Mahmood et al., Nanoscale 10, 18323–18330 (2018). 5. C. Spägele et al., Nat. Commun. 12, 3787 (2021). 6. R. C. Devlin, A. Ambrosio, N. A. Rubin, J. P. B. Mueller, F. Capasso, Science 358, 896–901 (2017). 7. N. A. Rubin et al., Science 365, eaax1839 (2019). 8. Y. Intaravanne, X. Chen, Nanophotonics 9, 1003–1014 (2020). 9. Y. Deng et al., Adv. Mater. 30, e1802632 (2018). 10. K. Huang et al., Laser Photonics Rev. 13, 1800289 (2019). 11. C. Zhang et al., Light Sci. Appl. 9, 55 (2020). 12. H. Liu et al., Nat. Phys. 14, 1006–1010 (2018). 13. A. Ahmadivand et al., Nano Lett. 19, 605–611 (2019). 14. M. Semmlinger et al., Nano Lett. 19, 8972–8978 (2019). 15. M. L. Tseng et al., Sci. Adv. 8, eabn5644 (2022). 16. W. K. Tobiska, A. Nusinov, in 35th COSPAR Scientific Assembly (Committee on Space Research, 2004), pp. 1–10. 17. F. Krausz, M. Ivanov, Rev. Mod. Phys. 81, 163–234 (2009). 18. B. Turkot, S. Carson, A. Lio, in 2017 IEEE International Electron Devices Meeting (IEDM) (IEEE, 2017), pp. 14.4.1–14.4.3. 19. Materials and methods are available as supplementary materials. 20. A. Thompson et al., “X-ray data booklet” (LBNL/PUB 490 Rev. 3, Lawrence Berkeley National Laboratory, 2009). 21. D. L. Windt, Comput. Phys. 12, 360–370 (1998). 22. E. D. Palik, G. Ghosh, Eds., Handbook of Optical Constants of Solids (Academic Press, 1998). 23. S. W. D. Lim, M. L. Meretska, F. Capasso, Nano Lett. 21, 8642–8649 (2021). 24. M. Hentschel et al., Light Sci. Appl. 12, 3 (2023). 25. C. M. Smith et al., Nature 424, 657–659 (2003). 26. T. W. Ebbesen, H. J. Lezec, H. F. Ghaemi, T. Thio, P. A. Wolff, Nature 391, 667–669 (1998). 27. J.-S. Park et al., Nano Lett. 19, 8673–8682 (2019). 28. W. T. Chen et al., Nat. Nanotechnol. 13, 220–226 (2018). 29. S. M. Kamali, E. Arbabi, A. Arbabi, Y. Horie, A. Faraon, Laser Photonics Rev. 10, 1002–1008 (2016). 30. P. B. Corkum, Phys. Rev. Lett. 71, 1994–1997 (1993). 31. M. Lewenstein, P. Balcou, M. Y. Ivanov, A. L’Huillier, P. B. Corkum, Phys. Rev. A 49, 2117–2132 (1994). 32. A. L’Huillier, K. J. Schafer, K. C. Kulander, J. Phys. At. Mol. Opt. Phys. 24, 3315–3341 (1991). 33. A. H. Firester, M. E. Heller, P. Sheng, Appl. Opt. 16, 1971–1974 (1977). 34. R. L. Lucke, “Rayleigh-Sommerfeld diffraction vs Fresnel-Kirchhoff, Fourier propagation, and Poisson’s spot” (NRL/FR/7218-04- 10,001, Naval Research Laboratory, 2004). 35. B. E. A. Saleh, M. C. Teich, Fundamentals of Photonics (Wiley, 2007). 36. H.-S. Chon et al., J. Opt. Soc. Am. A Opt. Image Sci. Vis. 24, 60–67 (2007). 37. M. Ossiander et al., Nat. Commun. 13, 1620 (2022). 38. M. Rimmer, Appl. Opt. 9, 533–537 (1970). 39. J. Zhu et al., Chin. Phys. B 29, 047501 (2020). 40. G. W. Webb, I. V. Minin, O. V. Minin, IEEE Antennas Propag. Mag. 53, 77–94 (2011). 41. M. Ossiander et al., figshare (2023); https://doi.org/10.6084/ m9.figshare.22191457. AC KNOWLED GME NTS Funding: This work was performed, in part, at the Center for Nanoscale Systems (CNS), a member of the National Nanotechnology Coordinated Infrastructure (NNCI), which is supported by the NSF under award no. ECCS-2025158. CNS is a part of Harvard University. The computations in this paper were run on the FASRC Cannon cluster supported by the FAS Division of Science Research Computing Group at Harvard University. M.O. acknowledges funding by the Alexander von Humboldt Foundation (Feodor-Lynen Fellowship), the Austrian Science Fund (FWF, Start Grant Y1525), and the European Union (grant agreement 101076933 EUVORAM). The views and opinions expressed are, however, those of the author(s) only and do not necessarily reflect those of the European Union or the European Research Council Executive Agency. Neither the European Union nor the granting authority can be held responsible for them. S.W.D.L. is supported by A*STAR Singapore through the National Science Scholarship Scheme. F.C. acknowledges financial support from the Air Force Office of Scientific Research (AFOSR) under award number FA9550-21-1- 0312. Author contributions: M.O. developed the project, designed the metalenses, and conducted the numerical modeling. M.L.M. fabricated the samples. M.L.M., S.W.D.L., and M.O. imaged the samples. H.K.H., M.O., N.K., T.J., and M.S. designed, conducted, and analyzed the experiment. M.O., M.S., and F.C. wrote the manuscript. All authors discussed the final version of the manuscript. Competing interests: M.O., M.L.M., S.W.D.L., and F.C. have filed a provisional patent application (US 63/385,066). The authors declare no other competing interests. Data and materials availability: The data that support the findings of this study are included in the manuscript and the supplementary materials and are archived at figshare (41). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/ science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg6881 Materials and Methods Figs. S1 to S6 References (42–62) Submitted 15 January 2023; accepted 7 March 2023 10.1126/science.adg6881 Ossiander et al., Science 380, 59–63 (2023) 7 April 2023 5 of 5
10.1126_science.adg7883
RES EARCH STRUCTURAL BIOLOGY Structure of the R2 non-LTR retrotransposon initiating target-primed reverse transcription Max E. Wilkinson1,2,3,4,5, Chris J. Frangieh1,2,3,4,5,6, Rhiannon K. Macrae1,2,3,4,5, Feng Zhang1,2,3,4,5* Non–long terminal repeat (non-LTR) retrotransposons, or long interspersed nuclear elements (LINEs), are an abundant class of eukaryotic transposons that insert into genomes by target-primed reverse transcription (TPRT). During TPRT, a target DNA sequence is nicked and primes reverse transcription of the retrotransposon RNA. Here, we report the cryo–electron microscopy structure of the Bombyx mori R2 non-LTR retrotransposon initiating TPRT at its ribosomal DNA target. The target DNA sequence is unwound at the insertion site and recognized by an upstream motif. An extension of the reverse transcriptase (RT) domain recognizes the retrotransposon RNA and guides the 3′ end into the RT active site to template reverse transcription. We used Cas9 to retarget R2 in vitro to non-native sequences, suggesting future use as a reprogrammable RNA-based gene-insertion tool. N on–long terminal repeat (non-LTR) ret- rotransposons are the most abundant class of mobile genetic element (MGE) in the human genome, primarily rep- resented by the LINE-1 and SINE (or Alu) long and short interspersed nuclear ele- ments, respectively (1). Despite their prevalence and contribution to genetic diversity and dys- regulation through mutagenicity and recombi- nation (1–3) and their prospective use as gene insertion tools, there is much left to understand about the mobility mechanisms of non-LTR retrotransposons (4). Pioneering research on the Bombyx mori (silk moth) R2 element (R2Bm), which selectively inserts into the 28S ribosomal RNA (rRNA) gene, has contributed substantially to our understanding of this type of MGE (5). R2, like all non-LTR retrotranspo- sons, encodes an open reading frame (ORF) with DNA binding, endonuclease, and reverse transcriptase activities (Fig. 1A). The endonu- clease domain (restriction-like endonuclease, RLE) nicks the target DNA, and the reverse transcriptase domain uses the exposed 3′ end from the nick to prime reverse transcription of the R2 RNA, resulting in a new genomic copy of the R2 element (Fig. 1B) (6, 7). This process is called target-primed reverse tran- scription (TPRT) and is characteristic of non- LTR retrotransposons and their group II intron ancestors (8, 9). The nicked strand that primes reverse transcription is referred to as the bot- tom strand. Complementarity between the bot- tom strand and the 3′ end of the R2 RNA (3′ 1Howard Hughes Medical Institute, Cambridge, MA 02139, USA. 2Broad Institute of MIT and Harvard, Cambridge, MA 02142, USA. 3McGovern Institute for Brain Research, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 4Department of Brain and Cognitive Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 5Department of Biological Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 6Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. *Corresponding author. Email: zhang@broadinstitute.org homology) is not required to initiate reverse transcription (10). Non-LTR retrotransposons are specific for reverse transcribing their own RNA; for R2, this specificity requires an ele- ment in the 3′ untranslated region (3′UTR), but the precise motif has not been located (11). It is also unclear how R2 specifically recog- nizes the 28S rRNA target gene, or how DNA nicking is coupled to reverse transcription within the same protein. To address these questions, we solved a cryo–electron micros- copy (cryo-EM) structure of R2Bm initiat- ing TPRT at the 28S rRNA gene using its own 3′UTR. The structure reveals an extensive in- terface with the target DNA, a small core re- gion of the 3′UTR required for TPRT, and shows that R2Bm can be engineered to reprogram its insertion site. Reconstitution and cryo-EM structure of an R2 TPRT complex We overexpressed R2Bm in Escherichia coli and purified it to apparent homogeneity (fig. S1). The purified protein was active in vitro, reproducing previously found biochemical ac- tivities, including RNA-stimulated nicking of the target DNA bottom strand, site-specific TPRT when supplied with in vitro–transcribed 3′UTR RNA, and low levels of template jump- ing (Fig. 1C) (6, 12). It is unclear if 3′ homology is required for TPRT in vivo; however, con- sistent with previous findings, we found that downstream sequences of up to 10 nucleotides (nt) do not inhibit activity in vitro (Fig. 1C) (10). Sequencing of TPRT junctions confirmed that homology-mediated TPRT is more likely to initiate reverse transcription at the 3′ end of the 3′UTR rather than skipping bases or inserting untemplated nucleotides (fig. S2) (10). To assemble a complex stalled during initiation of TPRT, we incubated R2Bm with target DNA, 3′UTR RNA, and the chain-terminator nucle- otide 2′,3′-dideoxythymidine (ddT), which mi- mics the first nucleotide incorporated in the TPRT reaction (dT) but does not allow further elongation. Purified TPRT complexes contained stoichiometric amounts of R2Bm, 3′UTR RNA, and target DNA with >99% of the bottom strand nicked (fig. S1). Initial attempts at cryo- EM imaging failed owing to the preferred orientation and flexibility of the complex. To overcome these issues, we used a carbon sup- port on the cryo-EM grid and added 5 nt of downstream 28S rRNA sequence to the 3′ end of the 3′UTR RNA to stabilize the complex by forming a primer-template duplex with the target DNA bottom strand. With these mod- ifications, we obtained a cryo-EM reconstruc- tion of the R2 TPRT complex at 3.1-Å resolution (Fig. 1D, figs. S3 and S4, and table S1). The core of the R2Bm protein is a reverse- transcriptase (RT) domain similar to that of group II intron RTs (13), followed by a C-terminal a-helical thumb domain and preceded by a characteristic N-terminal extension domain (NTE0) implicated in template switching (14), but the R2Bm RT includes a further N-terminal extension (NTE-1) that binds the 3′UTR RNA (Fig. 1, E and F) (15). Preceding the NTE-1 element are two DNA binding domains: the N-terminal C2H2 zinc finger domain (N-ZnF) and a Myb domain. C-terminal to the thumb domain lies an a-helical linker domain that packs against the thumb, followed by a CCHC zinc-finger domain (ZnF) conserved in many LINE ORFs (4). The ZnF then links to the C-terminal RLE domain, which cleaves the target DNA. This domain arrangement closely resembles that of Prp8 (13, 16, 17), the core protein of the spliceosome, underscoring the close relationship between pre-mRNA splicing and retrotransposons. There are several key interactions between the R2Bm protein, 3′UTR RNA, and target DNA (Fig. 1, E and F). The two strands of the target DNA separate around the ZnF domain, with the bottom strand feeding into the RLE active site where the scissile phosphate remains bound, while the top strand snakes along the opposing surface of the RLE. The RT active site contains a heteroduplex formed by the nicked bottom strand of the target DNA (5′ to the cleavage site) and the 5 nt of 28S rRNA homology extension beyond the 3′UTR RNA (Fig. 1G). This target heteroduplex is sur- rounded by residues important for RT activity (18), and the cryo-EM density shows incorpo- ration of the ddT chain terminator nucleotide into the bottom strand (Fig. 1H). The 5′ end of the bottom strand remains base-paired to the top strand as it leaves the RLE, and this down- stream DNA region has weak cryo-EM den- sity, suggesting that it is not tightly bound by R2Bm. The 248-nt 3′UTR RNA is mostly not resolved in the cryo-EM density except for a core 40-nt region, which wraps around the NTE-1 a helix of R2Bm and the 3′ end of which is guided into the RT active site via the NTE0 domain. Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Cryo-EM structure of the R2Bm retrotransposon. (A) Domains of the R2Bm retrotransposon. ZnF, zinc finger; NTE, N-terminal extension; RT, reverse transcriptase; RLE, restriction-like endonuclease. (B) Schematic of target-primed reverse transcription (TPRT). (C) Denaturing gel of in vitro TPRT reactions on a labeled 211-bp 28S DNA target. The same gel was visualized by Cy5 fluorescence and toluidine blue staining. (D) Cryo-EM density of the R2Bm TPRT complex. (E) Cartoon of the cryo-EM structure. Stars represent active sites. (F) Atomic model for the R2Bm TPRT complex. (G) Reverse transcriptase domain and template– primer duplex. (H) Reverse transcriptase active site. Cryo-EM density is shown as a gray transparent surface. R2Bm recognizes a sequence motif upstream of the cleavage site The target 28S DNA sequence has extensive interactions with R2Bm (summarized in Fig. 2A). Upstream bases from –38 to –7 and down- stream bases from +6 to +21 are respectively paired, whereas the 11 base pairs from –6 to +5 are melted around the RLE domain (bases are numbered relative to the bottom-strand cleavage site). The upstream DNA has a 40° bend and binds along the surface of the RT, linker, and thumb domains in a manner sim- ilar to that of the DNA in a recent group IIC intron maturase structure (Fig. 2B and figs. S5 and S6) (19). Many of the contacts between Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E A B D F R958 scissile +1 A T U A cDNA T A A ... +5 G G C T C G A C 3 +5 +1 248 RT C WT: 1: 2: 3: 4: -34 RUM -22 RASIN -27 DNA: WT 1 2 3 4 -35 -30 -25 -20 -15 -10 -7 R198 K149 R151 H673 K675 R119 H130 D761 K844 5 top strand C G G G T A A A C G G C G G G A G T A A C T A T G A C T C T C RUM bottom strand G C C C A T T T G C C G C C C T C A T T G A T A C T G A G A G 3 N1103 R935 A913 R919 C 1 2 6 A K964 R 7 7 9 R1101 T1100 +1 +5 RLE T A G C C A G G R 9 0 1 Q905 T T P 9 1 5 RASIN -5 R 9 2 4 R 9 2 2 A A T T C C D902 -1 R1093 L1090 N1098 G957 G1094 +10 A A A T G C C T C T T T A C G G A G 3 5 T189 R188 R125 K676 R132 R133 R517 H677 K772 R771 K786 R876 Q949 S1014 R961 V1012 3 base contact stacking Myb RT6a loop N-ZnF ZnF RLE downstream DNA +20 +10 3 5 RLE active site RASIN bottom strand Zn ZnF 1 6 Linker ~120o RLE active site RASIN top strand RLE -6 -1 NTE0 RT Myb Zn N-ZnF RT6a RUM N-ZnF NTE-1 upstream DNA RT RNA DNA bottom strand DNA top strand upstream DNA DNA top strand DNA bottom strand RUM library RASIN Nt.BbvCI RUM-specific cleavage adapter ligation, amplification, sequencing R2Bm G N-ZnF Zn R125 A( A( A( T( G( K149 Myb RUM R198 A( G( RUM H G( RUM top strand C( G( K675 H673 RT6a downstream DNA R2Bm: R2Bm: WT 6a = Cy5 (bottom strand) 28S RUM: RUM screen: 2 s 1 t i b 0 E I C G AA A T A T T A A G T T A C T C A GC C C T G A A T G C 34 22 position relative to bottom strand cleavage 27 T G A 25 A T A T 30 -34 RUM -22 RUM-RASIN distance -6RASIN 0 0 +3 +2 +1 R2Bm: = Cy5 (bottom strand) Fig. 2. Target DNA recognition upstream of the R2 cleavage site. (A) Schematic of interactions with the target DNA. Bases are numbered relative to the bottom- strand cleavage site. Positions of protein domains are shown by shaded rectangles. (B) Structure of R2Bm around the upstream DNA sequences. (C) Effect of upstream DNA mutations on target cleavage. The schematic shows the sequences of five DNA sequences tested in top-strand sense; dots represent bases identical to those of wild type. Red triangle, bottom-strand cleavage site. Denaturing gels show in vitro TPRT reactions on labeled 211-bp 28S DNA targets. DN, deletion of N-terminal N-ZnF and Myb domains. DRT6a, deletion of residues 672 to 677 (DGHRKK) of the RT6a loop. (D) Screen for identifying active RUM sequences. Nicking sites of R2Bm and the restriction endonuclease Nt.BbvCI are shown by triangles. (E) Sequence logo for sequences enriched in the RUM screen. (F to H) Details of interactions between the target DNA and the N-ZnF, Myb, and RT6a loop. (I) Effect of altering the distance between the RUM and RASIN motifs. Denaturing gel shows in vitro TPRT reactions on labeled 211-bp 28S DNA targets. Single-letter abbreviations for the amino acid residues are as follows: A, Ala; D, Asp; E, Glu; F, Phe; H, His; K, Lys; L, Leu; P, Pro; Q, Gln; R, Arg; S, Ser; T, Thr; V, Val; and Y, Tyr. Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Target DNA recognition at the R2 cleavage site. (A) Interactions of the top and bottom strands of the target DNA with the ZnF domain of R2Bm. Star, RLE active site. (B) Interactions of the DNA bottom strand with the RLE domain. (C) Interactions of the DNA top strand with the RLE domain. Residues mutated in the RD>AA mutant are highlighted. (D) RASIN sequence requirements for bottom- strand cleavage. The labeled 211-bp 28S DNA targets were incubated with R2Bm and 3′UTR RNA in the absence of deoxynucleotide triphosphates (dNTPs). The reactions were analyzed with a denaturing gel. Mutations are notated in top-strand sense, but both strands were mutated. (E) Denaturing gel showing R2Bm cleavage and TPRT activity on partially stranded substrates. Reactions contained a fluorescein-labeled 76-nt bottom strand. Reactions as indicated also contained 17 nt of downstream top-strand sequence (17d), 32 nt of upstream top strand sequence (32u), or 60 nt of top-strand sequence fully complementary to the bottom strand spanning the upstream and downstream regions. RD>AA, R2Bm R901A D902A. R2Bm and the DNA are via the phosphate backbone, suggesting that they are not se- quence specific. Based on the structure, how- ever, we predicted that two regions are key for sequence-specific DNA recognition by R2Bm: a 13-bp upstream motif from –34 to –22, which is bound by the N-terminal N-ZnF and Myb domains, and the 7 bp from –6 to +1, which are bound by the RLE (Fig. 2A). We term these re- gions the retrotransposon upstream motif (RUM) and retrotransposon-associated inser- tion site (RASIN), respectively. Consistent with the importance of the RUM region for R2 activity, mutating the entire up- stream sequence between –38 to –7 eliminated bottom-strand cleavage, whereas mutating the downstream sequences between +6 and +37 preserved wild-type levels of bottom-strand cleavage and TPRT (Fig. 2C) (20). Adding just the 13-bp RUM region to the upstream mutant at positions –34 to –22 restored near–wild- type activity, whereas a point mutant RUM (G–27 to C) did not rescue activity (Fig. 2C). This region of the target was strongly pro- tected in a previous deoxyribonuclease (DNase) footprinting assay (21). To systematically de- termine the importance of each base within the RUM, we performed an R2 cleavage assay on a DNA target with the upstream region (–38 to –7) mutated and the RUM (–34 to –22) replaced with a 13N library (Fig. 2D). Sequenc- ing of cleaved targets revealed a consensus RUM sequence A–31WWWGCNNNA–22, where W is A/T and N is any nucleotide, with minor preferences in other positions (Fig. 2E). This consensus is a close match to the wild-type 28S sequence A–31ACGGCGGGA–22, with the differences shown in bold. The RUM is recognized by three domains: N-ZnF, Myb, and an R2-specific insertion “6a” in the RT domain between motifs 6 and 7 (Fig. 2B and fig. S7). The N-ZnF has the classical C2H2 fold with a zinc ion coordinated between an a helix and a b hairpin, but unusually the a helix binds in the widened minor groove of the DNA from bases –18 to –23 instead of the typ- ical major groove (Fig. 2F and fig. S6) (22). The preference for A at base –22 in the RUM is likely due to N-ZnF Arg125, which hydrogen bonds with the minor-groove–facing side of the A–T base pair (Fig. 2F). The Myb domain forms a typical three-helix bundle, with the third helix bound in the major groove from bases –31 to –34 (22) while its linker to N-ZnF engages with base –30 (Fig. 2G). This is remi- niscent of other Myb–DNA structures, in- cluding telomere-interacting protein Rap1 (23). The Myb domain recognizes the A at base –31 through hydrogen bonds with Lys149 (Fig. 2G). Although Arg198 contacts bases at positions –33 and –34, these contacts appear not to be sequence specific, as the RUM screen showed only weak sequence preferences in this region (Fig. 2, E and G). Deletion of the N-ZnF and Myb domains together (DN mutant) complete- ly inhibits target DNA nicking and subsequent TPRT (Fig. 2C) (20). The central GC of the RUM is recognized by His673 and Lys675 of the loop 6a of the RT domain (Fig. 2H). Structural pre- dictions suggest that this loop is specific among non-LTR RT domains to R2 proteins (fig. S7). We found that deletion of the 6a loop inhibits Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Interactions of R2Bm with the 3′UTR RNA. (A) Secondary-structure diagram of the 3′UTR RNA, based on (26). Thicker strokes represent nucleotides visible in the cryo-EM density. Nucleotides are numbered from the first base of the 3′UTR (the base following the stop codon). (B) Structure of the 3′UTR RNA core and the R2Bm NTE-1 domain. Dotted lines, hydrogen bonds. (C) Low-pass filtered cryo-EM map. (D) Interactions between 3′UTR bases. Dotted lines, hydrogen bonds. (E) Secondary structure of the R2 tag RNA. Unshaded bases are not in the full-length 3′UTR. (F) Denaturing gel of in vitro TPRT reactions on a labeled 211-bp 28S DNA target using various R2 RNAs. Highlighted mutants are in the J1/2 region. The same gel was visualized by Cy5 fluorescence and toluidine blue staining. (G) The R2-tag allows TPRT of cargo RNAs. Denaturing gel shows TPRT reactions with equimolar amounts of the indicated RNAs and a labeled 211-bp 28S DNA target. R2 tag (43 nt) was added to the 3′ end of a 239-nt RNA encoding the CMV promoter or a 764-nt RNA encoding GFP. target DNA nicking (Fig. 2C). Finally, we found that the distance between the RUM and the bottom-strand cleavage site (the RASIN) is im- portant. Increasing the distance by one base was tolerated, but further increase or any de- crease to the distance inhibited target cleavage (Fig. 2I). Target DNA interactions at the cleavage and integration site The second key region for DNA target recog- nition by R2Bm is the target site for nicking by the RLE domain and R2 insertion, which we term the RASIN. In our structure, the 11 bp of the RASIN from –6 to +5 are melted around the RLE domain. The ZnF appears to act as the “zip,” stacking on the last upstream pair C–G(–7) with Arg922 and Arg924 and holding unzipped strands apart (Fig. 3A). Strand melt- ing may be enhanced by the 40° bend in target DNA around the RUM (Fig. 1F). Bases –6 to –1 on the bottom strand then follow a cleft be- tween the ZnF and the RLE, which adopts a canonical PD-(D/E)xK-family nuclease fold, but with the characteristic Lys1026 on an a helix instead of the usual b strand (Fig. 3B) (24). This lysine, along with catalytic residues Asp996 and Asp1009, is 4 to 6 Å from the scissile phosphate of C(–1), suggesting that C(–1) may be close to its position during catalysis of bottom-strand cleavage. On the top strand, bases –6 to +2 all make extensive contacts along a cleft between the RLE and linker domains, except for A(–4), which flips out and contacts C126 of the 3′UTR (Fig. 3C). To determine the relative impor- tance of the bases in the RASIN, we mutated each of the 11 bp individually and tested the effect on bottom-strand cleavage. Mutating T(+1) to A abolished cleavage entirely, and mutating T(–6), T(–5), and A(–3) severely de- creased activity, whereas other changes were tolerated (Fig. 3D). This suggests the following RASIN motif for cleavage, given in top-strand sense: T–6TNANNT+1. Because only the bottom strand of the RASIN enters the RLE active site, we tested the ac- tivity of R2Bm on a single-stranded DNA with the bottom-strand sequence and found that it was cut, albeit weakly (Fig. 3E). Endonuclease activity was strongly stimulated by providing a 60-nt top strand spanning the RASIN and up- stream and downstream sequences, but was similarly stimulated by a 32-nt top strand com- plementary only to the upstream region contain- ing the RUM. A 17-nt top strand complementary to the downstream sequence did not stimulate activity (Fig. 3E). This suggests that the RUM in a double-stranded state is important for re- cruiting the R2Bm RLE to the RASIN bottom strand and that the top strand of the RASIN, despite its extensive interaction with R2Bm, is dispensable for specific bottom-strand cleav- age. However, when we added deoxynucleotides Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. Mechanism and retargeting of first-strand synthesis by R2Bm. (A) Model for the initial stages of target site cleavage and first-strand synthesis. (B) Design of R2Bm + Cas9 experiments. (C) Complementation of DNA target site mutants by Cas9 cleavage in trans and cis. The denaturing gel shows in vitro TPRT reactions on a labeled 211-bp target corresponding to the wild-type 28S target, or two 235-bp targets: one where the RASIN TAAGGTA is replaced by 31 bp of unrelated sequence, and another where the 13-bp RUM is additionally scrambled. R2Bm and SpCas9(H840A) were added in trans, or in cis connected by a 33XTEN linker (fusion indicated by a shaded box). The sgRNA is complementary to the inserted sequence and nicks 40 nt from the last RUM base. The R2 RNA is the 3′UTR with 5 nt of 3′ homology to the nick site. (D) Sequences used for retargeting R2Bm to an unrelated locus from the Drosophila virilis genome. (E) Denaturing gel of in vitro TPRT reactions on the labeled 192-bp Drosophila virilis target. sgRNAs are numbered as in (D); all R2 RNAs or R2-tagged RNAs have 10 nt of 3′ homology to the nick site of the sgRNA. to these reactions, TPRT activity was eliminated in the absence of the top strand from the RASIN downstream but was partially rescued if the 3′UTR RNA contained 3′ homology to the tar- get site (Fig. 3E). The top-strand RASIN bases A(–4), A(–3), and G(–2) are grasped by Arg901 and Asp902 of the R2Bm linker (Fig. 3C). We mutated these two residues to alanine and tested TPRT activity on a fully double-stranded substrate, and found that TPRT activity was reduced and partially rescued by 3′ homology (Fig. 3E). These results suggest two important factors for initiating TPRT when the 3′UTR RNA lacks 3′ homology: (i) the presence of a top strand downstream of the RASIN, which may help retain the nicked bottom strand, and (ii) contacts between R2Bm and the top-strand RASIN, which help the nicked bottom strand “pivot” into the RT active site. R2Bm binds a small core region of the 3'UTR R2Bm can only initiate TPRT on RNAs contain- ing the R2 3′UTR (self-specificity), but the mo- lecular basis for this is not known (25). Multiple Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E models have been proposed for the secondary structure of the R2 3′UTR, and the divergent sequences of R2 RNAs have hindered identi- fication of key bases (26, 27). A model for the R2 3′UTR secondary structure based on chem- ical probing is shown in Fig. 4A and has at least 11 stems (26). In our cryo-EM map, we resolved density for two stems and their flank- ing single-stranded regions (Fig. 4B). On the basis of nomenclature commonly used for struc- tured RNAs, we name these stems P1 (nu- cleotides 33 to 38 and 120 to 135) and P2 (nucleotides 131 to 137 and 236 to 242), and term the single-stranded junction between P1 and P2 as J1/2 and the single-stranded region preceding P1 as J0/1. The rest of the 3′UTR may occupy a diffuse cloud of cryo-EM density next to these core regions (Fig. 4C). P1 and J1/2 are mainly recognized by an a helix from the R2Bm NTE-1 domain, which packs into the major groove of P1 and is wrap- ped by J1/2 (Fig. 4B). Arg307 recognizes the Hoogsteen edge of P1 G33, and the interaction is secured by Arg310 and Arg311. Consistently, these residues were previously shown to be essential for RNA binding (15), and the first 45 bases of the 3′UTR are essential for TPRT activity (11). J1/2 makes numerous sequence- specific contacts (Fig. 4D): A127 forms a sugar- edge pair with the Watson-Crick face of J0/1 A32, A128 hydrogen bonds to Leu732 and Lys733 of the R2Bm thumb domain and stacks on NTE-1 Tyr314, U129 hydrogen bonds to Glu319 and Lys322 of NTE-1, and C126 stacks on and hydrogen bonds with A(-4) from the top strand of the DNA target (Fig. 4, B and D). To test if regions of the R2 3′UTR not clearly visible in the cryo-EM density are required for TPRT activity, we designed a 43-nt minimal 3′UTR—“R2 tag”—that contains only the se- quences visible in the cryo-EM density, linked by tetraloops (Fig. 4E). The R2 tag was reverse transcribed as efficiently as the full 248-nt 3′UTR in a TPRT reaction. We tested the impor- tance of the J1/2 linker by making single-base transversions and found that A127U reduced activity and A128U almost completely abol- ished TPRT activity (Fig. 4F). Mutating G33 to C to disrupt base pairing at the bottom of stem P1 also reduced activity but could be rescued by the compensatory C125G mutation (Fig. 4F). Mutation of J0/1 A32 to G reduced activity, but mutations to C or U were tolerated. Equiv- alents to P1, P2, J0/1, and J1/2 can be identi- fied in the secondary structures of diverse R2 elements (26) (fig. S7). The P1 and P2 stems have different sizes and base compo- sitions, but positions 2 and 3 of J1/2, cor- responding to A127 and A128, are conserved as adenosines, consistent with their importance for TPRT. Because the R2 tag alone is efficiently in- tegrated in a TPRT reaction, we tested if ad- ding the R2 tag to the 3′ end of a “cargo” RNA would allow its integration at the 28S target site. We added the R2 tag to the 3′ end of a 239-nt cytomegalovirus (CMV) promoter RNA. This tagged RNA was used as efficiently as wild-type R2 3′UTR in a TPRT reaction, whereas an untagged RNA was not used, nor was an RNA tagged with an R2-tag A128U mutant (Fig. 4G). A larger RNA containing the 720-nt coding sequence for green fluorescent protein (GFP) and a 3′ R2 tag was also reverse tran- scribed in a TPRT reaction (Fig. 4G). R2Bm can be retargeted with CRISPR-Cas9 Our structural and biochemical observations suggest a multistep model for initiation of TPRT: The R2Bm N-terminal domains first detect a RUM sequence, followed by cleav- age of the bottom strand at the RASIN site, possible pivoting of the nick around the top strand into the RT active site, annealing of any 3′ homology to the nicked bottom strand, and finally initiation of reverse transcription (Fig. 5A). This model implies that R2Bm could prime reverse transcription off an exogenously nicked bottom strand close to the R2Bm bind- ing site (Fig. 5B). To test this, we replaced the RASIN and downstream sequences of the 28S DNA target with an unrelated sequence con- taining an efficient SpCas9 target sequence, but kept the RUM sequence to anchor R2Bm (Fig. 5B). This substrate could not be cleaved by R2Bm but was nicked efficiently by an SpCas9 H840A nickase mutant (Fig. 5C). When SpCas9 and R2Bm were added together with a single- guide RNA (sgRNA) and an R2 3′UTR RNA with 5 nt of 3′ homology to the sgRNA nick site, we detected low amounts of TPRT activ- ity. This activity was enhanced when the R2Bm and SpCas9 proteins were fused with a 33XTEN flexible linker (Fig. 5C). The RUM was not re- quired for Cas9-directed TPRT, as mutating the RUM did not reduce activity (Fig. 5C). This suggests that Cas9 might be able to direct R2Bm to perform TPRT at loci other than the 28S target. We mixed the R2Bm-Cas9(H840A) fusion protein with a 192-bp target sequence from Drosophila virilis, various sgRNAs, and R2 3′UTRs with 10 nt of 3′ homology to the nick site dictated by the sgRNA (Fig. 5D). We found TPRT activity at all Cas9 nick sites, with one sgRNA (guide 2) giving efficient activity (Fig. 5E). Adding R2Bm and SpCas9(H840A) as separate polypeptides also yielded efficient TPRT with guide 2 but was less robust with other guides (fig. S9). The 239-nt CMV promo- ter RNA with a 3′ R2 tag and 10 nt of homo- logy to the guide 2 nick site was also reverse transcribed efficiently; this activity required the R2 tag and was reduced in the absence of 3′ homology or with the R2 tag A128U muta- tion (Fig. 5E). Larger RNAs such as GFP could also be reverse transcribed at the guide 2 nick site (fig. S9). In summary, R2Bm can be retar- geted by Cas9 to perform TPRT at unrelated loci, and the R2 tag can direct incorporation of cargo RNAs at these sites. Discussion Here we show the structure of a non-LTR ret- rotransposon during transposition, and we dis- sect the principles of target DNA and self-RNA recognition. Our structure suggests that R2Bm uses its N-ZnF and Myb domains to locate the endonuclease target sequence, a model that contrasts with the model for other non-LTR retrotransposons in which the endonuclease domain is the only determinant of target site selection (28, 29). We identified two essential target site motifs—the RUM and RASIN—that are recognized by R2Bm, but we note that searching the B. mori genome with a RUM- RASIN consensus motif yields many potential off-target sites outside of the ribosomal DNA arrays (fig. S10). We examined the sequence of a previously identified B. mori non-28S inser- tion in (30) and found that the target site had limited similarity with 28S but had a TTAAcG|T RASIN motif (“|” indicates insertion site, low- ercase denotes deviation from 28S) and a GCTACTTGCGCAT RUM the correct distance upstream of the RASIN (fig. S10). Non-28S in- sertions, however, are rare, so it is likely that other factors are important in regulating R2Bm transposition, including chromatin accessibil- ity, other sequence motifs, or the ability of the target DNA to bend and melt. Non-LTR retrotransposons form a diverse family, and even within the R2 superclade there are notable differences between elements. R2Bm is a representative of the R2-D clade of elements, which have a single C2H2 N-terminal ZnF domain, but R2-A clade elements have three tandem N-terminal ZnF domains (31) that may create a more extensive DNA binding interface with greater stringency in target site selection. More broadly, non-LTR retrotrans- posons can be divided into two types on the basis of their endonuclease domains: those that, like R2Bm, use a C-terminal restriction enzyme-like (RLE) domain, and those that, like human LINE-1, use an unrelated N-terminal apurinic or apyrimidinic endonuclease (APE) domain (32, 33). Structure prediction using AlphaFold (34) suggests that, in these retro- transposons, the position of the APE domain is distinct from that of the RLE domain in R2Bm, suggesting that there may be mecha- nistic differences in how target cleavage is coupled to reverse transcription (fig. S5) (35). Nonetheless, the similarity between the DNA interface on the R2Bm thumb domain and the corresponding interface in the group IIC in- tron (fig. S5) suggests that this interface might be conserved among most non-LTR retrotrans- posons (19). Indeed, the upstream DNA from R2Bm was easily modeled into an AlphaFold model of human LINE-1 ORF2, including not only the thumb interactions but also strand Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E separation by the CCHC ZnF domain, which in LINE-1 ORF2 corresponds to the C-terminal cysteine-rich domain (fig. S5). Overall, the results of this work advance our understanding of transposition by non-LTR retrotransposons and suggest avenues for en- gineering these transposons for targeted gene insertions. RE FE RENCES AND N OT ES 1. H. H. Kazazian Jr., J. V. Moran, N. Engl. J. Med. 377, 361–370 (2017). 2. S. J. Priest et al., Nat. Microbiol. 7, 1239–1251 (2022). 3. S. R. Richardson et al., Microbiol. Spectr. 3, MDNA3–0061– 2014 (2015). 4. H. Fujiwara, Microbiol. Spectr. 3, MDNA3–0001–2014 (2015). 5. T. H. Eickbush, D. G. Eickbush, Microbiol. Spectr. 3, MDNA3–0011–2014 (2015). 6. D. D. Luan, M. H. Korman, J. L. Jakubczak, T. H. Eickbush, 7. Cell 72, 595–605 (1993). J. Yang, H. S. Malik, T. H. Eickbush, Proc. Natl. Acad. Sci. U.S.A. 96, 7847–7852 (1999). 8. A. M. Lambowitz, S. Zimmerly, Cold Spring Harb. Perspect. Biol. 3, a003616 (2011). 9. S. Zimmerly, H. Guo, P. S. Perlman, A. M. Lambowitz, Cell 82, 545–554 (1995). 10. D. D. Luan, T. H. Eickbush, Mol. Cell. Biol. 16, 4726–4734 (1996). 11. D. D. Luan, T. H. Eickbush, Mol. Cell. Biol. 15, 3882–3891 (1995). 12. A. Bibiłło, T. H. Eickbush, J. Mol. Biol. 316, 459–473 (2002). 13. C. Zhao, A. M. Pyle, Nat. Struct. Mol. Biol. 23, 558–565 (2016). 14. A. M. Lentzsch, J. L. Stamos, J. Yao, R. Russell, A. M. Lambowitz, J. Biol. Chem. 297, 100971 (2021). 15. V. K. Jamburuthugoda, T. H. Eickbush, Nucleic Acids Res. 42, 8405–8415 (2014). 16. W. P. Galej, C. Oubridge, A. J. Newman, K. Nagai, Nature 493, 638–643 (2013). 17. M. M. Mahbub, S. M. Chowdhury, S. M. Christensen, Mob. DNA 8, 16 (2017). 18. S. C. Pimentel, H. E. Upton, K. Collins, J. Biol. Chem. 298, 101624 (2022). 19. K. Chung et al., Science 378, 627–634 (2022). 20. S. M. Christensen, A. Bibillo, T. H. Eickbush, Nucleic Acids Res. 33, 6461–6468 (2005). 21. S. Christensen, T. H. Eickbush, J. Mol. Biol. 336, 1035–1045 (2004). 22. A. Klug, Annu. Rev. Biochem. 79, 213–231 (2010). 23. P. König, R. Giraldo, L. Chapman, D. Rhodes, Cell 85, 125–136 (1996). 24. A. Govindaraju, J. D. Cortez, B. Reveal, S. M. Christensen, Nucleic Acids Res. 44, 3276–3287 (2016). 25. M. Osanai, H. Takahashi, K. K. Kojima, M. Hamada, H. Fujiwara, Mol. Cell. Biol. 24, 7902–7913 (2004). 26. A. M. Ruschak et al., RNA 10, 978–987 (2004). 27. D. H. Mathews, A. R. Banerjee, D. D. Luan, T. H. Eickbush, D. H. Turner, RNA 3, 1–16 (1997). 28. H. Takahashi, H. Fujiwara, EMBO J. 21, 408–417 (2002). 29. Q. Feng, J. V. Moran, H. H. Kazazian Jr., J. D. Boeke, Cell 87, 905–916 (1996). 30. Y. Xiong, W. D. Burke, J. L. Jakubczak, T. H. Eickbush, Nucleic Acids Res. 16, 10561–10573 (1988). 31. A. Luchetti, B. Mantovani, PLOS ONE 8, e57076 (2013). 32. H. S. Malik, W. D. Burke, T. H. Eickbush, Mol. Biol. Evol. 16, 793–805 (1999). 33. I. R. Arkhipova, Mob. DNA 8, 19 (2017). 34. J. Jumper et al., Nature 596, 583–589 (2021). 35. I. Miller et al., Nucleic Acids Res. 49, 11350–11366 (2021). ACKN OWLED GMEN TS We thank S. Zhu and M. Walsh for valuable discussions; E. Brignole and C. Borsa for the smooth running of the MIT.nano cryo-EM facility, established in part with financial support from the Arnold and Mabel Beckman Foundation; S. Lövestam for a critical reading of the manuscript; and the entire Zhang lab for support and advice. We thank T. H. Eickbush and colleagues for their inspiring and pioneering work on R2 elements. Funding: Helen Hay Whitney Foundation Postdoctoral Fellowship (M.E.W.), Howard Hughes Medical Institute (M.E.W., F.Z.), National Institutes of Health grant 2R01HG009761-05 (F.Z.), Hock E. Tan and K. Lisa Yang Center for Autism Research at MIT (F.Z.), K. Lisa Yang and Hock E. Tan Molecular Therapeutics Center at McGovern (F.Z.), K. Lisa Yang Brain-Body Center at MIT (F.Z.), Broad Institute Programmable Therapeutics Gift Donors (F.Z.), the Pershing Square Foundation, W. Ackman, and N. Oxman (F.Z.), Asness Family Foundation (F.Z.), BT Charitable Foundation (F.Z.), the Phillips family (F.Z.), J. and P. Poitras (F.Z.), D. Cheng (F.Z.), R. Metcalfe (F.Z.). Author contributions: M.E.W. and F.Z. conceived the project. M.E.W. designed and performed experiments and solved the cryo-EM structure. C.J.F. generated and analyzed sequencing data. F.Z. supervised the research and experimental design with support from R.K.M. M.E.W. wrote the manuscript with input from all authors. Competing interests: F.Z. is a scientific adviser for and cofounder of Editas Medicine, Beam Therapeutics, Pairwise Plants, Arbor Biotechnologies, and Aera Therapeutics. F.Z. is a scientific adviser for Octant. Data and materials availability: The cryo-EM map has been deposited in the Electron Microscopy Data Bank with accession code EMD- 40033. The coordinates for the atomic model have been deposited in the Protein Data Bank with accession code 8GH6. The raw cryo-EM data have been deposited in EMPIAR with accession code EMPIAR-11458. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/ science-licenses-journal-article-reuse. This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the author-accepted manuscript of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg7883 Materials and Methods Figs. S1 to S10 Tables S1 to S3 References (36–47) MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. Submitted 20 January 2023; accepted 21 March 2023 10.1126/science.adg7883 Wilkinson et al., Science 380, 301–308 (2023) 21 April 2023 8 of 8
10.1126_science.adh0059
RES EARCH MOLECULAR BIOLOGY Structural basis for inactivation of PRC2 by G-quadruplex RNA Jiarui Song1,2,3, Anne R. Gooding1,2,3, Wayne O. Hemphill1,2,3, Brittney D. Love4,5,6, Anne Robertson4,5,6,7, Liqi Yao1, Leonard I. Zon4,5,6,7, Trista E. North4,5,6, Vignesh Kasinath1*, Thomas R. Cech1,2,3* Polycomb repressive complex 2 (PRC2) silences genes through trimethylation of histone H3K27. PRC2 associates with numerous precursor messenger RNAs (pre-mRNAs) and long noncoding RNAs (lncRNAs) with a binding preference for G-quadruplex RNA. In this work, we present a 3.3-Å-resolution cryo–electron microscopy structure of PRC2 bound to a G-quadruplex RNA. Notably, RNA mediates the dimerization of PRC2 by binding both protomers and inducing a protein interface composed of two copies of the catalytic subunit EZH2, thereby blocking nucleosome DNA interaction and histone H3 tail accessibility. Furthermore, an RNA-binding loop of EZH2 facilitates the handoff between RNA and DNA, another activity implicated in PRC2 regulation by RNA. We identified a gain-of-function mutation in this loop that activates PRC2 in zebrafish. Our results reveal mechanisms for RNA-mediated regulation of a chromatin-modifying enzyme. M any nuclear proteins that bind chro- matin also bind RNA molecules (1–3). The binding of RNA has been sug- gested to facilitate both positive and negative regulation (e.g., recruitment to target sites and enzymatic inhibition, re- spectively). Polycomb repressive complex 2 (PRC2) is a prominent example of a chroma- tin modifier known to be regulated by RNA (4, 5). PRC2 is essential for embryonic develop- ment and cell differentiation (6, 7). Some tu- mors are PRC2 dependent (e.g., because of silencing of tumor suppressor genes), making PRC2 a target for cancer therapeutics (8). PRC2 consists of four core protein components: EZH2 (catalytic subunit), EED [binds histone H3 tri- methylated at lysine 27 (H3K27me3)], SUZ12 (provides a platform), and RBAP48 (7). Asso- ciating cofactors define two PRC2 subclasses (9, 10), of which PRC2.2, containing AEBP2 and JARID2, is the subject of this study. PRC2 binds numerous pre-mRNA and long noncoding RNA (lncRNA) transcripts in vitro and in vivo (11–13). This broad RNA recogni- tion can be explained at least in part by PRC2 preferring an RNA G-quadruplex (G4) motif (14–16), which could be ubiquitous in the tran- scriptome from intramolecular and perhaps even intermolecular assemblies (17). Proposed models of RNA regulation of PRC2 remain disparate. First, in the “handoff” model, PRC2 requires RNA for recruitment and occupancy 1Department of Biochemistry, University of Colorado Boulder, Boulder, CO 80303, USA. 2BioFrontiers Institute, University of Colorado Boulder, Boulder, CO 80303, USA. 3Howard Hughes Medical Institute, University of Colorado Boulder, Boulder, CO 80303, USA. 4Stem Cell and Regenerative Biology Department, Harvard University, Cambridge, MA 02138, USA. 5Stem Cell Program, Division of Hematology/Oncology, Boston Children’s Hospital and Dana-Farber Cancer Institute, Boston, MA 02115, USA. 6Harvard Medical School, Boston, MA 02115, USA. 7Howard Hughes Medical Institute, Harvard Medical School, Boston, MA 02115, USA. *Corresponding author. Email: thomas.cech@colorado.edu (T.R.C.); vignesh@colorado.edu (V.K.) on a specific subset of targeted chromatin (18, 19). The direct handoff from RNA to DNA is an intrinsic property of PRC2, as shown by recent biophysical analyses (20). Second, the “eviction” model suggests that nascent RNA removes PRC2 from actively transcribed chro- matin to restrict nonspecific activity (14, 21–23). Third, in the “inhibitor” model, RNA and nu- cleosome binding of PRC2 are mutually exclu- sive, so RNA serves as a direct competitor to prevent PRC2 action (14, 18, 22, 24). Another version of the inhibitor model proposes that RNA exploits a regulatory site on PRC2 to abol- ish H3K27me3 binding to EED, which conse- quently eliminates allosteric activation of EZH2 (25). Therefore, structural details of PRC2-RNA interaction have been needed in the field to provide mechanistic insights and coordinate those models. Cryo–electron microscopy (cryo-EM) and x-ray crystallography have provided visualization of both substrate-free and nucleosome-bound PRC2 complexes (26–34). However, solving a structure of a PRC2-RNA complex has proved challenging. A streptavidin-biotin–affinity EM grid approach has been successfully used in cryo-EM (35), and here we adapt this technique for ribonucleoprotein (RNP) complexes using biotinylated RNA. We found that PRC2 can dimerize following RNA binding with a protein- protein interface composed of EZH2 CXC domains. The structure provides a molecular explanation for how RNA acts as a PRC2 in- hibitor, and it suggests a mechanism for RNA facilitation of PRC2 recruitment. Structure of G-quadruplex RNA-mediated PRC2 dimer We prepared six-subunit PRC2.2 complexes with full-length EZH2, SUZ12, RBAP48, EED, em- bryonic short-isoform AEBP2, and truncated JARID2119–450 (Fig. 1A), as well as two G4-forming RNAs (Fig. 1B) that bind PRC2 in vitro (fig. S1). We used streptavidin-affinity EM grids to cap- ture 5′-biotinylated RNAs, which select for RNA-bound PRC2 complexes and also protect them from the hydrophobic water-air inter- face. Because this method had not been ap- plied to an RNP complex, we validated it by testing different RNA concentrations with the same excess of PRC2. The number of particles observed by negative staining was proportional to the RNA concentration in most fields (Fig. 1C), indicating that the vast majority of protein complexes on the grid are bound to RNA. Com- pared with RNA-free PRC2 (particles ~150 Å in diameter), the majority of two-dimensional (2D) class averages from 1G4- and 2G4-bound PRC2 complexes were larger (particles ~250 Å in diameter), containing two recognizable PRC2 complexes (Fig. 1D). We determined the cryo-EM structure of the dimeric PRC2-1G4 RNA complex at 3.4-Å res- olution from consensus refinement, and at 3.3 Å from multibody refinement (36) (figs. S2 and S3 and table S1). The two PRC2 protomers in the RNP complex are nearly identical and have a conformation previously characterized as the SANT1 extended form (37) (Fig. 1E and fig. S4). We identified a protein interface in the RNA-induced dimer that is a localized EZH2- EZH2 interaction (described in the dimer in- terface section below) (movie S1). Notably, this PRC2 dimer has imperfect C2 symmetry (fig. S5A) associated with differ- ential occupancy of RNA in the two symmet- ric sites (fig. S5B). The stronger density has a volume representative of a G4 RNA (fig. S5, C and D, and S6D), whereas the symmetric site has discontinuous density and is not discussed hereafter. We could not obtain high-quality RNA density for de novo modeling in either of the sites from multibody refinement (fig. S5, C and D) and particle subtraction classification (fig. S6) (38). We attribute this to the multiple independent and flexible interactions between PRC2 and RNA. The G4 RNA is not nestled into the surface of the protein, as is typically seen for RNA-protein complexes, but instead appears to be separated from the protein. The RNA density is in closest proximity to the EZH2 SANT2 domain (residues 353 to 362), EZH2 CXC domain (residue R532), EZH2 SET domain (residue N697), the RRM (RNA recognition motif)–like domain of SUZ12, and an unstruc- tured region of RBAP48 (residues 92 to 107) (movie S1). These binding sites, excluding RBAP48, are supported by previous in vitro and in vivo studies (25, 26, 39, 40). We were able to observe density that links regions proximal to PRC2 and G4 RNA at 3.9-Å resolution from particle subtraction and classification (de- scribed in the arginine-rich site section below). The distance between the two PRC2 proto- mers is reduced from 51 to 48 Å on the side with stronger G4 density (Fig. 1E, bottom), which not only explains the imperfect symmetry, but Song et al., Science 381, 1331–1337 (2023) 22 September 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Overall structure of a PRC2-1G4 RNP complex. (A) (Left) Schematic of the proteins in the PRC2.2 six-subunit complex. Transparent N-terminal region of JARID2 was not included. (Right) Coomassie-stained gel of purified PRC2. (B) The two RNA oligonucleotides used in this study. (C) Negative-staining EM images of streptavidin-affinity grids with excess PRC2 and various 1G4 RNA concentrations. (D) Negative-staining EM provided 2D-class averages of PRC2 alone collected from continuous carbon grids and PRC2-G4 RNAs from streptavidin-affinity grids. (E) (Top) Cryo-EM density map of PRC2 bound to 1G4 RNA. EZH2 (CXC-SET) of protomer 1 in blue, EZH2 (CXC-SET) of protomer 2 in light blue, and 1G4 RNA in orange. (Bottom) The atomic model of PRC2 bound to 1G4 RNA. Distances between EZH2 (SANT2) and SUZ12 (RRM-like) are highlighted by black arrows to emphasize the change from 1G4-binding. Song et al., Science 381, 1331–1337 (2023) 22 September 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. G4 RNA induces PRC2 dimerization in solution. (Left) Size-exclusion chromatography of PRC2 preincubated with 1G4, 2G4, and mock (protein only). Abs, absorbance; mAU, milli–absorbance unit. (Right) Standard curves were used to estimate the molecular weights of PRC2 complexes. also supports the model of a single G4 being sufficient for PRC2 dimerization. Although the stoichiometry of most RNPs is 2 PRC2:1 RNA in our preparations, we do not reject the pos- sibility of a PRC2 dimer engaging two inde- pendent G4 RNAs simultaneously. G-quadruplex RNA induces PRC2 dimerization in solution To validate the dependence of PRC2 dimeriza- tion on G4 RNA binding in solution, we used analytical size-exclusion chromatography and mass photometry. In the absence of RNA, our six-subunit PRC2 complex chromatographed as a monomer (Fig. 2), which was consistent with an absolute molecular weight of 340 kDa (fig. S7A). Incubating PRC2 with 18-kDa 1G4 RNA or 30-kDa 2G4 RNA led to a large RNP complex of approximately 720 kDa measured by both size-exclusion chromatography (Fig. 2) and mass photometry (fig. S7C), which was consistent with a dimer. In addition, native gel electrophoresis of PRC2-G4 RNP cross-linked with glutaraldehyde indicated that G4 RNA re- mains bound as part of a cross-linked complex (fig. S7D). Furthermore, negative-staining EM reference-free 2D class averages of cross-linked complexes on conventional carbon-support grids confirmed the presence of RNA-mediated PRC2 dimers as observed in the non–cross- linked streptavidin-affinity grids (fig. S7E). Together, our results verified the requirement of RNA for this specific PRC2 dimerization in solution and provided confidence that the dimer was not an artifact of streptavidin-affinity selection. To test the specificity of G4 structure in me- diating PRC2 dimerization, we used microscale thermophoresis (MST) to measure PRC2-RNA binding in different reaction conditions (fig. S8). In a G4-favoring KCl buffer, the MST profile showed two distinguishable stages of thermo- phoretic mobility. This biphasic binding curve is typical for two binding events (41), which suggests that a higher-affinity binding site on PRC2 is primarily occupied at low PRC2 con- centrations (1 PRC2:1 RNA), and at higher PRC2 concentrations, lower-affinity binding of a sec- ond PRC2 follows (2 PRC2:1 RNA). In addition, we performed the same MST assays in a G4- destabilizing LiCl buffer or using a G-rich single-stranded RNA with no G4-forming po- tential. Neither experiment gave a distinct bi- phasic curve, indicating that PRC2 dimerizes specifically on RNA containing at least one G4 motif. The dimer interface prevents nucleosome and H3 tail binding The PRC2-1G4 RNA dimer has three features that would be expected to inhibit nucleosome binding and histone methylation. First, the residues within the CXC domain of EZH2 in- teract with the CXC of the second protomer to form the dimer interface. This dimer inter- face includes two R566–A569 hydrogen bonds, two R566–T573 hydrogen bonds, two Q575– G564 hydrogen bonds, and a K568–K568 hy- drophobic interaction (Fig. 3A). By contrast, in a nucleosome-bound PRC2, the CXC domain facilitates the catalytic activity of the adjacent SET domain of EZH2, specifically with R566, K568, T573, and Q575 contributing to interac- tions with nucleosome DNA and the H3 tail (Fig. 3B) (27). The disparate functions of the CXC domain in these different PRC2 structures are seen by the superposition of our density map onto the nucleosome-bound PRC2, which shows clashes with both the DNA and the H3 tail (Fig. 3C and movie S2). Therefore, we propose that nucleosome binding and H3 tail loading, both of which are essential for histone methy- transferase (HMTase) activity, are mutually antagonistic with RNA-mediated PRC2 di- merization. The disruption of nucleosome-PRC2 complexes by 1G4 RNA was confirmed by a competition-binding assay in solution (Fig. 3D). Second, the EZH2 bridge helix (residues 500 to 516)—important for nucleosome DNA bind- ing and channeling H3 tail into the active site of the EZH2 SET domain (27)—is disordered in both protomers of our RNP complex (fig. S9). This is consistent with structures of PRC2 lacking nucleosome substrate. Lastly, the EZH2 C-terminal helix (residues 738 to 742), which points away from the H3K27 binding site in nucleosome-bound PRC2, now occludes the active site in RNA-bound PRC2 (fig. S9) and appears to serve as an additional mechanism to prevent H3 tail binding. To test the importance of the CXC dimer interface, we purified a mutant (EZH2 R566A K568A Q575A). Mutation of these residues did not affect G4 RNA binding (fig. S10, A and B) or prevent RNA-induced dimerization (fig. S10, C and D). However, negative-staining EM with streptavidin-affinity grids classified sub- stantially more monomer-size particles from the mutant (59%) than the wild-type (WT) PRC2 (9%) (Fig. 3E and fig. S10E). This suggests that these EZH2 mutations impacted the overall stability of the PRC2 dimer by destabilizing the protein-protein interaction, and consequent- ly, one protomer more easily dissociated dur- ing stringent washes (dilutions) in our grid preparations compared with the WT. Because streptavidin-affinity grids only retain RNA-bound complexes, those monomer particles of the mu- tant could also represent an intermediate stage of one PRC2 engaging RNA prior to associa- tion of the second PRC2, which is consistent Song et al., Science 381, 1331–1337 (2023) 22 September 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. RNA-mediated PRC2 dimer is an inactive complex. (A) Cryo-EM structure of PRC2-1G4 complex with zoom-in to show the interface of two EZH2 CXC domains. Interacting residues (R566, K568, T573, and Q575) are highlighted in the stick representation. One set of hydrogen bonds (R566NE- T573OG1, 2.70-Å distance; R566NH1-A569O, 2.73-Å distance; and Q575NE2-G564O, 2.51-Å distance) is indicated by yellow dashed lines. The other set of hydrogen bonds between the same amino acid pairs in the second PRC2 protomer is not shown for clarity. Another view of the same region is shown in fig. S3G. (B) Structure of PRC2-nucleosome complex (PDB: 6WKR) with zoom-in to emphasize the CXC interactions with nucleosome H3 tail (orange). The same residues as shown in (A) are high- lighted in stick representation. (C) Superposition of the EZH2 CXC domain of the RNA-bound PRC2 on the nucleosome-bound PRC2 to empha- size disparate functions of the CXC domain. (D) Nucleosome-RNA competi- tion assay. PRC2 was incubated with constant amount of radiolabeled trinu- cleosome and serial dilutions of 1G4 RNA. Incomplete PRC2-trinucleosome complexes are indicated by * and **. We assume two of three nucleosomes were occupied by PRC2 in *, and one of three nucleosomes in **. (E) Negative-stain EM to quantify monomer and dimer particles of EZH2 R566A K568A Q575A binding 1G4. The number of particles in each class is indicated above each bar. (F) Binding affinity of 1G4 RNA or dsDNA to WT PRC2 and EZH2 R566Y K568Y W575Y (3Y) measured by FP. We used a reaction buffer with a lower salt concentration to achieve higher binding affinity for the dsDNA (right). Kd, dissociation constant. (G) Methyl- transferase activities from figs. S12 to S15 are plotted against reaction times for PRC2 (400 nM) preincubated with 1G4 (400 nM), 2G4 (400 nM), and mock (protein only). Error bars represent standard deviations of three replicates performed on different days. (H) Methylation of trinucleo- somes by PRC2 with serial dilutions of 1G4 RNA. (Top) Radiogram that shows methylation. (Bottom) Coomassie- stained gel. Single-letter abbreviations for the amino acid residues are as follows: A, Ala; C, Cys; D, Asp; E, Glu; F, Phe; G, Gly; H, His; I, Ile; K, Lys; L, Leu; M, Met; N, Asn; P, Pro; Q, Gln; R, Arg; S, Ser; T, Thr; V, Val; W, Trp; and Y, Tyr. Song et al., Science 381, 1331–1337 (2023) 22 September 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. EZH2 loops physically contact G4 RNA and contribute to direct handoff from RNA to DNA. (A) Map of PRC2-1G4 RNA from particle subtraction and classification (fig. S6) with zoom-ins to emphasize the observed physical interactions of PRC2 and G4 RNA. (B) EZH2 structure from AlphaFold predicts two disordered loops of EZH2. Arginine-rich loop [EZH2(353–362)] and lysine-rich loop [EZH2(494–502)] are indicated in blue and green, respectively. (C) FP assays to monitor the transfer kinetics of PRC2 from fluorescently labeled 1G4 RNA to a dsDNA competitor. The ratio kq/k–1 of the PRC2 RNA-to-dsDNA direct-transfer rate constant (kq) and the PRC2-RNA dissociation rate constant (k–1) provides a measure of the propensity of PRC2 to exchange these ligands by the direct transfer mechanism. WT PRC2 had kq = 90 ± 11 M−1s−1, k–1 = 5.6 ± 0.49 × 104 s−1, and kq/k–1 = 1.7 ± 0.35 × 105 M−1. EZH2 D353–362 had kq = 48 M−1s−1, k–1 = 12 × 104 s−1, and kq/k–1 = 0.4 × 105 M−1. EZH2 CR had kq = 130 ± 62 M−1s−1, k–1 = 2.7 ± 0.63 × 104 s−1, and kq/k–1 = 4.6 ± 1.3 × 105 M−1. koff obs, dissociation rate constant observed. with the biphasic binding curve observed in our MST assays. We next attempted to disrupt the CXC inter- face more severely by substituting bulky side- chains of tyrosine, so we constructed an EZH2 R566Y K568Y Q575Y mutant. Unexpectedly, this mutant had higher binding affinity to the G4 RNAs, as determined by fluorescence po- larization (FP) assays (Fig. 3F, left) and elec- trophoretic mobility-shift assays (EMSA) (fig. S11, A and B). Notably, double-stranded DNA (dsDNA) binding of this mutant was not af- fected (Fig. 3F, middle and right), which fur- ther indicates that DNA and RNA use separate mechanisms to engage PRC2 even though they bind mutually antagonistically. Although this mutant PRC2 is a monomer as observed by negative-staining EM, RNA-bound particles showed a dominant population of dimer com- plexes, which is consistent with the increased RNA binding affinity and the role of RNA in mediating PRC2 dimerization (fig. S11, C and D). We propose that the aromatic sidechains of tyrosine might stack on each other and therefore stabilize the CXC interface. Thus, the dimerization interface need not be spe- cific, and it appears to be RNA binding rather than protein-protein interaction that drives PRC2 dimerization. Overall, we observed a positive correlation between the CXC dimer interface and G4 RNA binding. RNA-induced PRC2 dimer is inactive Structural observations on the EZH2 CXC in- terface prompted us to hypothesize that the HMTase activity of the G4-induced PRC2 di- mer would be inhibited. To test this, we per- formed activity assays to compare free PRC2 with RNA-sequestered dimers (Fig. 3G and figs. S12 to S15). As expected, we detected sub- stantial reductions in methylation rates with all substrates (including recombinant H3) in response to RNA binding, with stronger inhibi- tion by the higher-affinity 2G4 RNA (Fig. 3G). The extent of inhibition was limited by the RNA concentration because complete inhibi- tion was achieved with excess 1G4 RNA (Fig. 3H and fig. S16A). An arginine-rich site of EZH2 binds G4 RNA and participates in RNA-to-DNA handoff Applying particle subtraction and classifica- tion (fig. S6), we identified multiple sites in EZH2 [EZH2(353–362): KRPGGRRRGR, EZH2 R532, and EZH2 N697] that physically contact G4 RNA (Fig. 4A). The arginine-rich EZH2(353– 362) has been implicated in binding lncRNA (25, 39), and similar arginine-rich sequences in multiple transcription factors have been linked to RNA binding (42). EZH2 truncation [EZH2(D353–362)] and a local charge-reversed EZH2 [EZH2 CR(353–362): DEPGGEEEGE] both exhibited decreased HMTase activity but showed no obvious reduction in G4 RNA binding or G4 RNA-mediated PRC2 dimerization in vitro (fig. S17). We also generated a double-truncation mutant [EZH2(D353–362 D494–502)] to remove an adjacent lysine-rich site (EZH2 494–502) (Fig. 4B). EZH2 residues 494 to 502 were previously implicated in binding nucleosome DNA and G4 RNA (27, 40) and might com- pensate for RNA binding in the absence of EZH2 residues 353 to 362. PRC2 containing Song et al., Science 381, 1331–1337 (2023) 22 September 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. EZH2 CR is a gain-of-function mutant in rescue of zebrafish develop- ment. (A) Representative images of injected zebrafish embryos at 48 hours post fertilization (hpf). Gross phenotypic scoring of anterior-posterior axis growth was sorted into three categories: normal, reduced, and severely reduced growth. (B) Scoring of anterior-posterior axis growth at 48 hpf. Zebrafish embryos were injected with 4-ng ezh2-MO or the same amount of control MO. For rescue experiments, 100 ng of mRNA encoding WT or mutated human EZH2 was coinjected with ezh2-MO. At least three clutches were examined for each injection. Fisher’s exact test was used to determine the P values. (C) Dose-response experiments with 25, 50, and 100 ng of mRNA coinjected. Statistical analyses were performed as in (B) by comparing different doses with ezh2-MO alone. ns, not significant; *P < 0.05; **P < 0.01; ***P < 0.001; ****P < 0.0001. EZH2(D353–362 D494–502) exhibited a 1.5- to twofold reduction of binding affinity for G4 RNA (fig. S17G). We attribute this modest re- duction to the presence of other RNA-binding regions observed in our structure and sup- ported by previous studies (25, 40, 43). PRC2 has the intrinsic ability to directly trans- fer or hand off from RNA to DNA, without there ever being a free-enzyme intermediate (20, 44). We therefore examined whether the arginine- rich EZH2(353–362) region, in addition to its role in RNA binding, is important for such RNA-to-DNA handoff. We found that the EZH2(D353–362) and EZH2 CR mutants had a 4.3-fold reduction and a 2.7-fold increase, respectively, in the propensity for direct trans- fer from RNA to DNA (Fig. 4C). We rationalized these results with a model (fig. S18) in which PRC2 harbors the arginine-rich EZH2(353–362) and lysine-rich EHZ2(494–502) to form a ternary intermediate with both RNA and DNA binding. EZH2(D353–362) and CR mutations of the RNA- binding region affect the propensity for direct transfer in different directions. EZH2 CR is a gain-of-function mutant in zebrafish development We used zebrafish to examine the significance of the EZH2 G4 RNA-binding sites in verte- brate development. Zebrafish and human EZH2 proteins have high sequence identity, including the regions responsible for G4 RNA binding and RNA-induced PRC2 dimerization (fig. S19A). EZH2 knockdown in zebrafish by antisense morpholino oligonucleotides (MOs) led to a severe growth defect (45) represented by gross alterations in the length of the anterior-posterior axis (Fig. 5A). As expected, coinjection with mRNA that encodes human WT EZH2 signif- icantly rescued the growth defects (P < 0.0001) (Fig. 5B and fig. S19B). Mutant EZH2 mRNAs rescued overall development to varying degrees (fig. S19C). Notably, the EZH2 CR mutant had a gain-of-function phenotype, giving significant- ly better rescue than that of WT EZH2 (P < 0.01) and phenotypically mimicking a gain-of- function mutant (EZH2 Y646F) that is well- studied in the human system and frequently found in lymphoma (Fig. 5B) (46–48). Dose- response assays by coinjecting ezh2-MO and in- creasing amounts of mRNAs (Fig. 5C) confirmed that the extent of rescue was consistently similar between CR and Y646F, whereas the catalyt- ically dead EZH2(D694–751) gave no rescue. Therefore, the EZH2 CR mutant, which shows an increased propensity for direct transfer be- tween RNA and DNA in vitro, also behaves as a gain-of-function mutant in zebrafish development. Discussion In the past 10 years, lncRNAs and pre-mRNAs have become prominent in discussions of PRC2 regulation (4, 5). Because of the broad PRC2 transcriptome (11, 12), deciphering molecular details of PRC2-RNA interaction has been chal- lenging. PRC2 binds G4 RNA in vitro (14, 15, 23), PRC2 binding sites on chromatin genome wide are associated with G-tract motifs (15), and a well-defined G4-forming RNA TERRA (telomeric repeat–containing RNA) recruits PRC2 to telomeres (16). Thus, we focused on G4 RNA in the present study. By using a biotinylated G4 RNA with streptavidin- biotin affinity grids, we determined the cryo-EM structure of an RNA-bound PRC2 complex. This structure supports the earlier conclusion that nucleosomal DNA and RNA binding are mutu- ally antagonistic (22, 24), but it provides a much more interesting mechanism than just competi- tion on overlapping sites. Instead, G4 RNA trig- gers formation of a PRC2 dimer that occludes the DNA-binding amino acids. Based on the pre- sent structure and biochemical and biophysical data, we propose a model that can explain multi- faceted functions of RNA in PRC2 regulation. First, actively transcribed loci, which need to avoid silencing by PRC2, generate nascent RNA transcripts that have the potential to di- merize PRC2 among other RNA processing events. The RNA-induced dimer simultaneously inactivates two PRC2 complexes (no H3 tail binding) and evicts them from local chromatin (no nucleosome DNA binding). Residues of the EZH2 CXC domain, known to load the H3 tail into the catalytic groove, are occupied in a protein-protein interaction that reinforces the dimerization. Dimerization may prevent the spreading of H3K27me3 across regions near preexisting H3K27me3 or JARID2 116me3 marks, both of which induce allosteric activa- tion of PRC2 (31, 49, 50), and therefore, it may help define heterochromatin boundaries. Song et al., Science 381, 1331–1337 (2023) 22 September 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E Second, the interactions of PRC2 with RNA are proposed to be important for PRC2 occu- pancy on chromatin (19, 51). Consistent with this idea, PRC2 has the intrinsic ability to translocate onto target chromatin as it simultaneously dis- sociates from inhibitory RNA (20). To reconcile the role of RNA in inhibiting and evicting PRC2 from chromatin and its role in promoting PRC2 chromatin occupancy, we propose that EZH2 harbors partially redundant nucleic acid bind- ing sites that allow PRC2 to transiently engage both RNA and DNA, thereby facilitating direct transfer from RNA to DNA. The EZH2 CR mu- tant, designed to destabilize RNA binding, en- hances the direct transfer of PRC2 from RNA to DNA in vitro. This CR mutant rescues the knockdown of PRC2 in zebrafish better than WT EZH2. This gain-of-function phenotype is consistent with a direct transfer from RNA to DNA, which facilitates PRC2 activity in vivo. However, the many differences between in vitro and in vivo experiments warrant a cautious in- terpretation. The precise mechanism of the EZH2 CR mutant gain-of-function merits fur- ther investigation. Our structure describes one mode of RNA recognition by PRC2, but there may very well be others. Other reported RNA-binding sites include the RNA-binding region (RBR) adja- cent to the bridge helix of EZH2 (residues 494 to 502) (40), the stimulatory recognition motif (SRM) of EZH2 (residues 127 to 153) (25), the EED amino acids close to EZH2 SRM (residues 336 to 355) (25), and the JARID2 RBR (residues 332 to 358) (43). Although we did not obtain any subclass map having a distinguishable RNA density in proximity to those regions, this is not sufficient to reject their RNA-binding potential. We propose that the numerous RNA-binding regions within PRC2 explain why mutations give only modest effects on RNA binding in this and other studies. PRC2 has been shown to dimerize without RNA binding. We consistently find that, in the absence of RNA, a small fraction of PRC2 mol- ecules self-associate into dimers at high protein concentrations, as found for a four-subunit PRC2 holoenzyme (52) and the six-subunit PRC2.2 complex (fig. S20A). However, 2D class averages of self-associated dimers show a dif- ferent dimer interface than the RNA-induced dimer (fig. S20B). In addition, two reported domain-swapped PRC2 dimers—PRC2-PCL and PRC2:EZH1 (29, 32)—have completely dif- ferent architectures than our RNA-mediated dimer. Aside from PRC2, many chromatin-associated complexes have been found to interact with RNA, including other histone modifiers (2), transcription factors (42, 53, 54), and DNA methyltransferase (55, 56). Similar to PRC2, they tend not to have canonical RNA-recognition motifs and bind RNA broadly, and obtaining molecular structures of the RNA-protein com- plexes has been very challenging. Ultimately, solving additional structures of RNA bound to epigenetic modifiers will reveal the mechanisms by which RNA serves direct regulatory roles, rather than simply serving as a messenger. R E F ER EN C E AN D N OT ES 1. D. G Hendrickson, D. R. Kelley, D. Tenen, B. Bernstein, J. L. Rinn, Genome Biol. 17, 28 (2016). 45. Y. Zhong, Q. Ye, C. Chen, M. Wang, H. Wang, Nucleic Acids Res. 46, 3382–3399 (2018). 46. C. J. Sneeringer et al., Proc. Natl. Acad. Sci. U.S.A. 107, 20980–20985 (2010). 47. D. B. Yap et al., Blood 117, 2451–2459 (2011). 48. G. P. Souroullas et al., Nat. Med. 22, 632–640 (2016). 49. S. Sanulli et al., Mol. Cell 57, 769–783 (2015). 50. Q. Zhang et al., Nat. Commun. 12, 4592 (2021). 51. A. Laugesen, J. W. Højfeldt, K. Helin, Mol. Cell 74, 8–18 (2019). 52. C. Davidovich, K. J. Goodrich, A. R. Gooding, T. R. Cech, Nucleic Acids Res. 42, 9236–9248 (2014). 53. R. Saldaña-Meyer et al., Mol. Cell 76, 412–422.e5 2. Y. Long, X. Wang, D. T. Youmans, T. R. Cech, Sci. Adv. 3, (2019). eaao2110 (2017). 3. R. Xiao et al., Cell 178, 107–121.e18 (2019). 4. C. Davidovich, T. R. Cech, RNA 21, 2007–2022 5. (2015). J. Yan, B. Dutta, Y. T. Hee, W. J. Chng, RNA Biol. 16, 176–184 (2019). 6. R. Cao et al., Science 298, 1039–1043 (2002). 7. R. Margueron, D. Reinberg, Nature 469, 343–349 (2011). 8. A. Laugesen, J. W. Højfeldt, K. Helin, Cold Spring Harb. Perspect. Med. 6, a026575 (2016). 9. S. Hauri et al., Cell Rep. 17, 583–595 (2016). 10. G. van Mierlo, G. J. C. Veenstra, M. Vermeulen, H. Marks, Trends Cell Biol. 29, 660–671 (2019). 11. J. Zhao et al., Mol. Cell 40, 939–953 (2010). 12. C. Davidovich, L. Zheng, K. J. Goodrich, T. R. Cech, Nat. Struct. Mol. Biol. 20, 1250–1257 (2013). 13. M. Rosenberg et al., Nat. Struct. Mol. Biol. 28, 103–117 (2021). 14. S. Kaneko, J. Son, R. Bonasio, S. S. Shen, D. Reinberg, Genes Dev. 28, 1983–1988 (2014). 15. X. Wang et al., Mol. Cell 65, 1056–1067.e5 (2017). 16. J. J. Montero et al., Nat. Commun. 9, 1548 (2018). 17. D. Varshney, J. Spiegel, K. Zyner, D. Tannahill, S. Balasubramanian, Nat. Rev. Mol. Cell Biol. 21, 459–474 (2020). 18. C. Cifuentes-Rojas, A. J. Hernandez, K. Sarma, J. T. Lee, Mol. Cell 55, 171–185 (2014). 19. Y. Long et al., Nat. Genet. 52, 931–938 (2020). 20. W. O. Hemphill, R. Fenske, A. R. Gooding, T. R. Cech, Proc. Natl. Acad. Sci. U.S.A. 120, e2220528120 (2023). 21. S. Kaneko, J. Son, S. S. Shen, D. Reinberg, R. Bonasio, Nat. Struct. Mol. Biol. 20, 1258–1264 (2013). 22. M. Beltran et al., Genome Res. 26, 896–907 (2016). 23. M. Beltran et al., Nat. Struct. Mol. Biol. 26, 899–909 (2019). 24. X. Wang et al., Nat. Struct. Mol. Biol. 24, 1028–1038 (2017). 25. Q. Zhang et al., Nat. Struct. Mol. Biol. 26, 237–247 (2019). 26. V. Kasinath et al., Science 359, 940–944 (2018). 27. V. Kasinath et al., Science 371, eabc3393 (2021). 28. S. Poepsel, V. Kasinath, E. Nogales, Nat. Struct. Mol. Biol. 25, 154–162 (2018). 29. D. Grau et al., Nat. Commun. 12, 714 (2021). 30. L. Gong et al., Nat. Commun. 13, 6781 (2022). 31. L. Jiao, X. Liu, Science 350, aac4383 (2015). 32. S. Chen, L. Jiao, X. Liu, X. Yang, X. Liu, Mol. Cell 77, 1265–1278.e7 (2020). 33. S. Chen, L. Jiao, M. Shubbar, X. Yang, X. Liu, Mol. Cell 69, 840–852.e5 (2018). 34. K. Finogenova et al., eLife 9, e61964 (2020). 35. B. G. Han et al., J. Struct. Biol. 195, 238–244 (2016). 36. T. Nakane, S. H. W. Scheres, Methods Mol. Biol. 2215, 145–160 (2021). 37. V. Kasinath, S. Poepsel, E. Nogales, Biochemistry 58, 346–354 (2019). 38. X. C. Bai, E. Rajendra, G. Yang, Y. Shi, S. H. Scheres, eLife 4, e11182 (2015). 39. S. Kaneko et al., Genes Dev. 24, 2615–2620 (2010). 40. Y. Long et al., eLife 6, e31558 (2017). 41. M. A. Lemmon et al., EMBO J. 16, 281–294 (1997). 42. O. Oksuz et al., Mol. Cell 83, 2449–2463.e13 (2023). 43. S. Kaneko et al., Mol. Cell 53, 290–300 (2014). 44. W. O. Hemphill, C. K. Voong, R. Fenske, J. A. Goodrich, T. R. Cech, Proc. Natl. Acad. Sci. U.S.A. 120, e2220537120 (2023). 54. H. R. Steiner, N. C. Lammer, R. T. Batey, D. S. Wuttke, Biochemistry 61, 2490–2494 (2022). 55. A. Di Ruscio et al., Nature 503, 371–376 (2013). 56. L. I. Jansson-Fritzberg et al., RNA 29, 346–360 (2023). AC KNOWLED GME NTS We thank R. Yan, Z. Yu, and S. Yang (Janelia Cryo-EM Facility); C. Moe (University of Colorado Boulder BioKEM Facility); and S. Zimmermann and G. Morgan (University of Colorado Boulder Electron Microscopy Service) for microscope operation and data collection. We thank A. Erbse (University of Colorado Boulder Shared Instruments Pool RRID SCR_018986) for mass photometer operation, data acquisition, providing the MST instrument, and help with MST data interpretation. We thank A. Iragavarapu for help with plasmid preparation, R. Fenske for help with FP assays, Y. Long (Weill Cornell Medicine) and C. Lim (University of Wisconsin Madison) for helpful discussions, and B. Greber for helpful suggestions related to model refinement. Funding: J.S. is supported by Howard Hughes Medical Institute (HHMI)–Jane Coffin Childs postdoctoral fellowship. W.O.H. is supported by NIH NIGMS postdoctoral fellowship (F32-GM147934). V.K. is supported by startup funds from University of Colorado Boulder and NIGMS R00GM132544. T.R.C. and L.I.Z. are HHMI Investigators. Author contributions: Conceptualization: J.S., V.K., and T.R.C. Investigation: all authors; Visualization: J.S., A.R.G., W.O.H., B.D.L., A.R., and L.Y.; Funding acquisition: L.I.Z., T.E.N., V.K., and T.R.C.; Project administration: V.K. and T.R.C.; Supervision: L.I.Z., T.E.N., V.K., and T.R.C.; Writing – original draft: J.S., V.K., and T.R.C.; Writing – review and editing: all authors. Competing interests: T.R.C. is a scientific advisor for Storm Therapeutics, Eikon Therapeutics, and Somalogic, Inc. L.I.Z. is a founder and stockholder of Fate Therapeutics, CAMP4 Therapeutics, Triveni Bio, Scholar Rock, and Branch Biosciences. L.I.Z. is a consultant for Celularity and Cellarity. The other authors declare no competing interests. Data and materials availability: All data are available in the main text or the supplementary materials. Cryo-EM density maps and fitted models have been deposited in the Electron Microscopy Data Bank (EMD-29578, consensus map; EMD-29647, Body1 from multibody refinement; and EMD-29656, Body2 from multibody refinement) and the Protein Data Bank (PDB: 8FYH). Requests for reagents, plasmids, cell lines, and zebrafish strains used in this study should be directed to the corresponding authors. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal-article- reuse. This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the Author Accepted Manuscript (AAM) of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh0059 Materials and Methods Figs. S1 to S20 Table S1 References (57–68) Movies S1 and S2 MDAR Reproducibility Checklist Submitted 3 February 2023; resubmitted 5 July 2023 Accepted 22 August 2023 10.1126/science.adh0059 Song et al., Science 381, 1331–1337 (2023) 22 September 2023 7 of 7
10.1126_science.adh5206
RES EARCH NEUROSCIENCE Volitional activation of remote place representations with a hippocampal brain–machine interface Chongxi Lai1*†‡, Shinsuke Tanaka1†‡, Timothy D. Harris1, Albert K. Lee1,2* The hippocampus is critical for recollecting and imagining experiences. This is believed to involve voluntarily drawing from hippocampal memory representations of people, events, and places, including maplike representations of familiar environments. However, whether representations in such “cognitive maps” can be volitionally accessed is unknown. We developed a brain–machine interface to test whether rats can do so by controlling their hippocampal activity in a flexible, goal-directed, and model-based manner. We found that rats can efficiently navigate or direct objects to arbitrary goal locations within a virtual reality arena solely by activating and sustaining appropriate hippocampal representations of remote places. This provides insight into the mechanisms underlying episodic memory recall, mental simulation and planning, and imagination and opens up possibilities for high-level neural prosthetics that use hippocampal representations. T he ability to simulate scenarios in one’s mind is a hallmark of intelligence, as it facilitates the evaluation of past exper- iences and future plans. For instance, we can imagine walking around our previ- ous workplace, or imagine how our current workplace might function if we rearranged the furniture. Such imagination requires an inter- nal world model that can be flexibly accessed to construct possible scenarios (1–3). The hippocampus is a brain region that is critical for memory and imagination (1, 4–6). It holds a model of the environment (also called a cognitive map) (7, 8) that could potentially be mentally traversed for the purpose of recall or simulation. In particular, the hippocampus contains spatial maplike representations of previously explored environments. Each envi- ronment’s representation consists of place cells—neurons that fire selectively whenever an animal moves through specific locations (called the “place fields” of those cells) in that environment (9, 10). This selective firing re- sults in a distinct multicell activity pattern at each location in the environment, which, during physical navigation, can be used to decode the animal’s current location from the ongoing pattern of neural activity (11). In contrast, a key aspect of imagination is the activation of neural representations that deviate from cur- rent sensory input, such as those that are non- local (i.e., represent locations away from one’s current location). Previous work has shown brief and intermittent activation of nonlocal 1Janelia Research Campus, Howard Hughes Medical Institute, Ashburn, VA, USA. 2Howard Hughes Medical Institute and Department of Neurology, Beth Israel Deaconess Medical Center, Boston, MA, USA. *Corresponding author. Email: chongxi.lai@gmail.com (C.L.); alee31@bidmc.harvard.edu (A.K.L.) †Present address: Howard Hughes Medical Institute and Department of Neurology, Beth Israel Deaconess Medical Center, Boston, MA, USA. ‡These authors contributed equally to this work. hippocampal spatial representations sugges- tive of the planning of specific paths within a cognitive map (12–21). However, it is unknown whether this activity is volitionally controlled or rather reflects passive memory–related processes that are presumably nonvolitional (22, 23). To test whether an animal can directly con- trol its hippocampal activity according to its model of the world, we used a brain–machine interface (BMI) approach because, unlike with humans, we cannot simply ask animals to imagine scenarios. With BMI methods, we could reward animals for generating neural activity resembling the simulation of specific scenar- ios. More precisely, we could reward them for the volitional activation of specific nonlocal representations from the cognitive map—a fundamental building block of scenario simu- lation. BMI research has a rich history of di- rectly testing for volitional control of activity patterns of neuronal ensembles in the motor cortex and related areas (24–35). In the hippo- campus, it has been shown that the activity level of individual neurons (36, 37) or the pop- ulation activity related to individual stimuli (38) can be controlled. However, a real-time BMI that allows humans or animals to con- trol their hippocampal population activity in terms of the content of their cognitive map (e.g., location representations) has never been demonstrated. A hippocampal map–based BMI We designed a real-time hippocampal BMI and two BMI tasks to investigate whether rats could navigate to goals (“jumper” navigation task), or move external objects to goals while remaining stationary (“Jedi” object location control task), within an immersive virtual real- ity (VR) environment solely by controlling the activity of a population of place cells. Each jumper or Jedi BMI experiment consisted of three phases (Fig. 1A). In phase 1, rats ran to a succession of arbitrary locations marked by a tall, visible goal cue placed in a familiar two- dimensional virtual arena (“running” task). Upon reaching each cue, liquid reward was delivered, the trial ended, and the cue moved to another location for the next trial. Animals were secured in a harness and could freely rotate their body and head direction on top of a spherical treadmill (39) while hippocampal CA1 neural activity was recorded (Fig. 1B, fig. S1, and movie S1). We applied a recently de- veloped field-programmable gate array (FPGA)– based neural signal processor to perform low- latency (1 ms) assignment of extracellular spikes (recorded from 128 channels) to a population of hippocampal units (40, 41). In the running task, treadmill movement updated the ani- mal’s location in the virtual environment, and many hippocampal units (i.e., place units) dis- played spatially modulated activity (39, 42–44) (Fig. 1B, blue arrows) similar to that in real- world environments (8–11). In phase 2, the binned spike counts from the most recent 1.5 or 5 s of activity of these place units and the animal trajectory from the running task were used to train a decoder (Fig. 1B, green arrows) that estimates the animal’s current location from the neural data every 100 ms. We used a deep neural network for decoding (fig. S2), allowing the use of data augmenta- tion for training—a method that improves both the decoder’s performance given limited data and its noise robustness. In phase 3, the treadmill was disconnected from the VR sys- tem, and the animal’s ability to control its own or an object’s translational movement was limited to controlling its hippocampal activity, which was converted by the decoder into a specific location output every 100 ms (Fig. 1C). Note that the decoder was trained to estimate the animal’s current location in the running task only, not its location in the subsequent BMI tasks, but, during BMI periods, the ani- mal needed to generate activity corresponding to locations away from its current location. BMI navigation task In the jumper task, we tested whether animals could navigate to arbitrary goal locations as in the running task, except here by means of BMI-based first-person teleportation. After rats performed the running task for ~40 min (~120 trials) (Fig. 2, A and B, and movie S1), the data were used to train the decoder, which accurately estimated the rat’s current location in the running task [validation set coefficient of determination (R2) = 0.78 to 0.88] (Fig. 2C). Jumper trials were identical to running trials, except the animal’s location was updated to the BMI-decoded location (smoothed with a 3-s sliding window to help reduce potential high-frequency visual jitter of the VR updates) (Fig. 2, D and E, and movie S1). If an animal did not reach the goal within 62 s, the trial Lai et al., Science 382, 566–573 (2023) 3 November 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E phase 1: Running task phase 2: Train decoder phase 3: BMI tasks A B R V e t a d p u Running task trial 1 trial 2 physically run to goals Train deep network population decoder of current location in Running task goal decoded location controlled object phase 1: Running task VR projector raw hippocampal neural data task 1: BMI navigation (“Jumper”) task: teleport self to goal (animal teleported toward decoded location) trial 1 trial 2 task 2: BMI object location control (“Jedi”) trial 1 trial 2 32 electrode groups 1 2 3-32 task: move object to goal & maintain it there (animal fixed at center, object moved toward decoded location) spike sort ~40 min data and select place units 360˚ screen spherical treadmill update location goal cue animal’s view of virtual arena s t i n u e c a p l treadmill movement phase 2: Train decoder of current location animal locations time 0 max firing rate spatially tuned activity of individual place units sliding window of spike counts 3 0 2 0 0 1 0 0 s t i n u e c a p l 0 0 0 2 1 0 0 time bins neural activity top view of virtual arena C phase 3: BMI tasks VR projector low-latency spike assignment using spike sorted model (in FPGA) goal cue real-time spike trains of place units R V e t a d p u spherical treadmill update location treadmill disconnected decoded location time real-time spike counts (in PC) 2 0 1 0 1 2 0 2 1 2 0 1 0 1 0 trained decoder Fig. 1. Hippocampal map–based brain–machine interface in a virtual reality system. (A) Steps for performing the two different BMI experiments in this study. Rats first physically ran to a series of goals (running task), while their hippocampal neural activity and (virtual) location in a square arena were recorded. This data was used to train a decoder to take neural activity as input and output the animal’s current location in the running task. In BMI task 1 (jumper task), animals needed to generate neural activity that would be decoded as locations they wanted to move to so that they could reach each goal (to obtain a reward). In BMI task 2 (Jedi task), animals were fixed at the center of the virtual arena (but could rotate) and needed to generate activity corresponding to locations where they wanted an external object to move to so that the object reached the goal, then they needed to sustain that activity to maintain the object there (to maximize reward). (B) Schematic of the VR system (left). The animal was free to rotate its body in the horizontal plane. In the running task, the animal’s location in the virtual arena environment was updated according to treadmill movement. Simultaneously recorded spiking from a population of hippocampal CA1 units expressed place fields—the basis of the cognitive map of the environment (right). Decoder was then trained using binned spiking activity and location data. (C) In both BMI tasks, the treadmill no longer updated VR. Instead, the animal or object location was controlled solely by real-time hippocampal activity. A neural signal processor rapidly assigned activity to individual units, whose spike counts were fed into the decoder. VR projection was updated according to locations output by the decoder. In the jumper (Jedi) task, the animal’s (object’s) virtual location was moved toward the most recent decoded locations. PC, personal computer. Lai et al., Science 382, 566–573 (2023) 3 November 2023 2 of 8 Phase 1: Running task trial 11 1m x 1m arena trial 10 goal B Running trial D Jumper BMI trial 20 cm 20 cm 6.4 s 3.4 s 3.5 s 4.2 s 7.6 s RES EARCH | R E S E A R C H A R T I C L E A trial end trial start 3.8 s trial 14 17.7 s 13.0 s 4.0 s 3.4 s 2.3 s 3.7 s 6.0 s trials 21-99 Phase 2: Train decoder of current location behavior decoded 20 cm e r o c s 10 s of trajectory Y X R² score = 0.78 (rat 1) m c 0 5 X Y 10 s 1 t a r 2 t a r 3 t a r ) m c ( r o r r e i g n d o c e d 40 20 0 rat 1 rat 2 rat 3 C F s t i n u e c a p l s / m c 50 0 V µ 0 0 2 56 32 8 ) z H ( y c n e u q e r f E Jumper BMI trial (no movement) ) e r o c s - z ( 5 y t i v i t c a t i n u -1 7 -1 treadmill speed 2 s LFP s / m c 50 0 2 s V µ 0 0 2 56 32 8 3 -1 e r o c s - z 150 0 0.5 0.0 s t i n u s / m c V µ 0 0 2 56 z H 32 8 treadmill speed 2 s 3 -1 3 -2 rat 2 p = 1.5 × 10-7 rat 3 p = 2.8 × 10-15 100 20 20 cm 5.4 s 125 25 Phase 3: “Jumper” BMI navigation task 1mx1m arena goal 10.6 s 10.3 s 10.0 s 11.0 s trial 1 trial 2 G rat 1 p = 8.2 × 10-11 control Jumper trial end trial start trial 5 7.6 s 8.0 s 16.6 s 5.8 s 6.6 s 11.1 s 14.2 s 9.0 s 120 t n u o c 20 30 14.6 mean trial duration (s) 50 40 16.3 30 40 50 14.3 30 40 50 H goal-directedness of trajectory 0° 0° 0° 45° -45° 45° -45° 45° -45° 11.6 s 7.2 s 12.2 s 90° -90° 90° -90° 90° -90° trials 13-53 Run Jumper 135° -135° 135° -135° 135° -135° 180° 180° 180° Fig. 2. Rats can navigate to goals by controlling their hippocampal activity. In both the running and jumper BMI tasks, animals were rewarded when they reached each goal. (A) Animal trajectories in the virtual arena for consecutive running task trials. Trial duration (time to reach goal) in seconds is shown. (B) Example running task trial. From top: trajectory, firing rate (z-scored) of individual units (units were ordered by time of peak activity), treadmill speed, and LFP from one recording channel and corresponding wavelet spectrogram during trial. (C) Accuracy of trained decoder of animal’s current location for held-out running task data. Actual and decoded trajectories during example trial (top left) and across several trials (for x and y coordinates separately, bottom left). Median decoding error (distance between actual and decoded locations) with range and quartiles (bottom right). (D) Example jumper BMI trial with similar trajectory as the running trial in (B). From top: trajectory generated by the animal controlling its hippocampal activity and the decoder output (animal is teleported toward decoded location; each gray circle represents the decoded location at the time the animal is at the corresponding point in the trajectory connected by the dark line, sampled here every 1 s), firing rate of individual units [using same order of units as in (B)], treadmill speed, LFP, and spectrogram. (E) Example jumper BMI trial in which animal did not move the treadmill. Trajectory (left) as in (D). Right, from top: unit activity, treadmill speed, LFP, and spectrogram. See fig. S10 for all 10 nonmovement trials. (F) BMI-generated trajectories for consecutive jumper trials. (G) Mean jumper trial duration (magenta vertical line) is significantly lower than distribution of expected mean duration for simulated trials if goals were in random locations. (H) Polar distribution of angle between direction of movement and direction to goal during running and jumper tasks. Zero corresponds to animal movement directly toward the goal center. Lai et al., Science 382, 566–573 (2023) 3 November 2023 3 of 8 1 t a r 2 t a r 3 t a r B s / s e k p s i 50 0 25 s / m c 0 50 m c 94599459945 0 32 s t i n u e c a p l 250 200 150 100 50 0 250 μV ) z H ( y c n e u q e r f 24 16 8 RES EARCH | R E S E A R C H A R T I C L E A Phase 1: Running task Phase 2: Train decoder of current location Phase 3: “Jedi” BMI object location control task 50 cm goal 5 s distribution of real-time decoded locations 0 max. density rat location (fixed at center) 950950950 9555595955 960960960 965965965 970970 975999975975 980980980 population burst event (PBE) population spike count treadmill speed distance to goal 0 time 5 s decode 20 cm 1 s rat location (fixed at center) 3 2 1 0 -1 e r o c s - z 250 200 150 100 50 0 250 μV 24 16 8 0 time 5 s 20 cm 1 s 1 s rat location (fixed at center) 3 2 1 0 -1 p=5.2 × 10-10 p=7.6 × 10-11 D 78.8% < 1 cm/s 67.4% < 1 cm/s 81.2% < 1 cm/s C p=1.8 × 10-22 control Jedi 100 t n u o c 20 100 20 100 20 14.8 30 40 19.1 25 30 35 18.0 25 30 35 mean distance to goal (cm) rat 1 rat 2 rat 3 F D C 1.0 0.8 0.6 0.4 0.2 0.0 1 cm/s 1.0 0.8 0.6 0.4 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 0 10 20 treadmill speed (cm/s) rat 1 0 10 20 30 0 5 10 15 rat 2 rat 3 Fig. 3. Rats can move objects to remote goal locations and maintain them there by controlling their hippocampal activity. In the Jedi BMI task, trials did not end when the external controlled object first reached the goal; instead, animals were rewarded as long as the object was in the goal region (white circle), for up to 3 min per trial. The animal was always fixed at center of the virtual arena but could rotate its body and generally turned toward each goal. (A) Distribution of real-time decoded locations (output every 100 ms) generated by the animal controlling its hippocampal activity across eight consecutive Jedi BMI trials for rats 1, 2, and 3. Panels show decoded locations during each trial (up to 3 min; fig. S11). Periods when the animal’s body rotated >12°/s were excluded. See text and methods for details. The external controlled object (which was visible for rats 1 and 2 but invisible for rat 3) was moved toward the decoded location (fig. S11 shows that the distribution of object locations was essentially the same as the distribution of decoded Lai et al., Science 382, 566–573 (2023) 3 November 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E locations). (B) A 40-s-long period during an example trial during which the animal did not move the treadmill. From top: summed activity across all units with PBEs identified, treadmill speed, distance of decoded location from goal (0 means inside goal region), and close-ups of two 5-s periods [as animal moves object to goal (left) and as animal maintains object at goal (right); points in the arena represent sequence of decoded locations] with spike trains of units, LFP, and spectrogram. See fig. S12 for additional example periods. (C) Mean distance of decoded location from goal across all trials (magenta vertical line) is significantly lower than mean distance expected for randomized goal locations. (D) Treadmill speed distribution during periods shown in (A), illustrating that the animal was generally still during task performance. ended and a new goal cue appeared at a ran- dom location. Rats successfully navigated by controlling their hippocampus, generating efficient paths to each goal (Fig. 2F; see figs. S3 to S5 for all trials of three rats, and figs. S6 to S8 for all trials re-decoded using a shorter decoding window and without smoothing). To check whether this performance could be attributed to non-spatially-specific neural activity (e.g., modulating global firing rate), we randomly shuffled the spike trains across place units, ran the shuffled data through the original decoder to produce simulated trajectories, then deter- mined how long it would have taken to reach the same sequence of goal locations as in the original experiment. Shuffled-unit mean trial durations were much longer than the actual means (P < 10−100, three rats, one session each), suggesting that performance depended on generating place field–related activity. To test whether generating non-goal-directed sequen- ces of location-specific activity (e.g., random movement within the cognitive map) could explain the performance, we randomly shuffled the goal locations in each trial while preserv- ing the original BMI trajectories and then de- termined the time that would have been needed to reach the shuffled goals. Shuffled- goal mean trial durations were again much longer than actual means (P = 2.8 × 10−15 to 1.5 × 10−7) (Fig. 2G), indicating that animals’ BMI trajectories were clearly goal-directed. Goal-directedness was also apparent from the distribution of angles between the animal’s instantaneous direction of BMI-generated movement and the direction from the animal’s current location to the goal, which was con- centrated around a value near 0° (Fig. 2H). Thus, even though jumper trials took longer than running trials (mean trial duration across animals: 15.1 s versus 6.9 s; note, however, that BMI decoding and smoothing added a few sec- onds to jumper durations), the animals’ routes revealed effective, goal-directed, map-based BMI navigation. Furthermore, such performance was achieved without extensive BMI training (Fig. 2F; figs. S3 to S5 show sessions 3, 9, and 2 for rats 1, 2, and 3, respectively; table S1; a fourth rat failed to perform either BMI task). Although animals were free to physically run during the jumper task, such movement was not necessary for task performance. Initially, animals ran as in the running task, but in later trials, animals ran less (fig. S9). In a subset of trials (10 out of 161 trials) (Fig. 2E and fig. S10), animals remained still, yet in all cases they efficiently reached the goal. Moreover, this suc- cessful navigation did not depend on activity in population burst events (PBEs), which often appear during immobility and during which brief activation of place cell representations for remote locations has been shown to occur (13–16, 18, 20, 21, 23). BMI object location control task Although episodic memories are encoded and often retrieved using a first-person perspective, individuals can also imagine scenarios from a third-person perspective, with other animate and inanimate players taking part. Further- more, imagination often involves holding a single thought in mind for extended periods. Therefore, our second BMI task, the Jedi task, tested whether animals—while remaining in the same place—could use the same map of the arena to control the location of a virtual object, guide it to the goal cue location, and maintain it nearby. The jumper and Jedi tasks thus used different forms of feedback: self- location and the location of an object, respec- tively. After the same running task and decoder training phases as in the jumper experiment, the animals in Jedi were fixed (but could freely turn) at the arena’s center, and the object’s location was updated to the BMI-decoded location (with a 2-s smoothing window). In each trial, the goal cue remained in the same place, providing re- ward as long as the object touched it. After 3 min or the rat having received 0.5 ml of reward in total, whichever came first, a new goal cue ap- peared at a distant random location for the next trial. Rats could activate and sustain a remote lo- cation’s representation around the goal for long periods, until the trial ended, and then shift attention to the next goal (Fig. 3, A and B; fig. S11; and movie S1). Performance was mea- sured using the mean distance (over time) be- tween the decoded locations and goals. Shuffling spike trains across units yielded much greater mean distances than the actual means (P = 2.2 × 10−5 to 2.6 × 10−3, three rats, one session each). To assess the goal-directedness of BMI- generated activity, we shuffled the goal loca- tions while preserving the locations output by the decoder. The decoded (and controlled object’s) location was far more concentrated around the actual remote goal cue than shuffled goal locations (P = 1.8 × 10−22 to 5.2 × 10−10) (Fig. 3C and fig. S11), indicating clear goal- directed control of activity. Again, such per- formance occurred without extensive training (Fig. 3A shows sessions 7, 6, and 3 for animals 1, 2, and 3, respectively). Task performance was not dependent on PBEs, as there was no change in performance when all activity in PBEs was eliminated and the decoder was rerun post hoc (fig. S11). Animal movement was generally low when engaged in the Jedi task (Fig. 3D), and move- ment was not required for successful perform- ance. There were many longer periods (≥8 s long with a treadmill speed of ≤1 cm/s, 38 periods, mean: 17.3 s, maximum: 44.0 s) during which the animal did not move the treadmill while it directed the object to the goal and/or held it there (34 of 38 periods) (Fig. 3B and fig. S12). Activity during PBEs was also generally not necessary for performance in these non- movement segments (fig. S12). Features of volitionally generated spiking and local field potential activity What characteristics did the volitionally gen- erated activity have? First, mean firing rates per unit were similar between jumper and run- ning tasks (fig. S13A). Mean firing rates per unit were correlated across Jedi and running tasks, but lower in Jedi (fig. S13B)—consistent with the decreased physical movement in Jedi. We then investigated the hypothesis that, to move themselves or the object toward a given (decoded) location in the jumper and Jedi tasks, animals generated a pattern of firing rates across units (i.e., a population vector, or PV) similar to the mean PV at that location over the entire running task (called the refer- ence PV, or rPV) (Fig. 4A). (Note that the set of rPVs for all locations is thus equivalent to the standard place field map across the popula- tion.) We examined the correlation between the PV generated at each moment (in every 500-ms window) during jumper or Jedi and the rPV of the decoded location at that mo- ment. As a benchmark, we computed the cor- relation between the 500-ms PVs during the running task with the rPVs corresponding to the animal’s actual location at those times (Fig. 4, B and C, “Run”) as well as the correla- tion between the running task PVs and the rPVs of random locations (Fig. 4, B and C, “randRun”). We then correlated jumper or Jedi PVs with the rPVs of the decoded locations at each moment (Fig. 4B for “jumper” and Fig. 4C and fig. S14 for “Jedi”) and with rPVs of ran- dom locations (“randJumper” and “randJedi”). Jumper and Jedi PVs were significantly correlated Lai et al., Science 382, 566–573 (2023) 3 November 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E A activity during Running task or BMI task s t i n u DNN decoder ^ ^ (x,y) B Jumper (T=500ms) 0.48 0.46 0.44 0.42 0.40 0.38 ) V P r , V P ( r r o c unit N T unit N 0.1 1.2 1.8 0.6 unit 1 0.3 0.0 0.7 2.2 0.1 0.1 corr(PV, rPV) PV (Hz) rPV (Hz) C Jedi (T=500ms) ) V P r , V P ( r r o c G unit 1 place fields (from Running task) 0.42 0.40 0.38 0.36 0.34 0.32 0.30 0.28 randRun Run randJumper Jumper randRun Run Jedi randJedi rat 1 rat 2 rat 3 0.45 0.40 0.35 0.30 0.25 0.28 0.26 0.24 0.22 0.20 0.18 0.35 0.30 0.25 0.20 0.15 0.250 0.225 0.200 0.175 0.150 0.125 0.100 D ] . d n a r - l a u t c a [ ) V P r , V P ( r r o c Δ 0.18 0.17 0.16 0.15 E 0.12 0.10 0.08 ] . d n a r - l a u t c a [ ) V P r , V P ( r r o c Δ 1.0 1.0 Jumper - randJumper Run - randRun 2.0 3.0 time window T (s) 4.0 5.0 Jedi - randJedi Run - randRun 2.0 3.0 time window T (s) 4.0 5.0 F DNN decoding window N i e s o n + V P s t i n u 1 0 (x,y) actual location rPV r DNN decoder Bayesian decoder ^ ^ (x,y) (x,y)^ ^ N s t i n u 1 i e s o n + V P r R2 score x m c 0 5 y m c 0 5 e r o c s 2 R 1.0 0.8 0.6 0.4 0.2 0.0 10s DNN Bayesian DNN R² = 0.84 Bayesian R² = 0.59 actual location rat 1 rat 2 rat 3 0 2.5 5 H rat location (fixed at center) goal decoded location density (without PBEs) 0 max ) z ( A U M s / m c m c 5 0 5 0 50 0 population burst event (PBE) population spike count mean i.i.d. noise level (Hz) 3s treadmill movement distance to goal (excluding PBEs) 100 ms LFP V µ 0 0 2 ) z H ( y c n e u q e r f 1.0 0.8 12 0.6 8 0.4 0.2 4 0.0 2.5 e r o c s - z -2 I ) z H / ² z ( l a n g s i d e r o c s - z f o D S P 0.08 0.06 0.04 0.02 0.00 no movement (Jedi) movement (Jedi) 5 10 15 20 25 frequency (Hz) Fig. 4. Volitionally generated nonlocal activity is similar to the activity when the animal is at the corresponding locations and is associated with theta-band power in the LFP. (A to E) The population vector (PV) of ongoing spiking activity was compared with the average place field activity (rPV) at a given location during the running task. (A) Schematic of comparison. (B) Mean correlation of instantaneous (500-ms window) PV during running or jumper task with rPV for the current location (in the running task), current decoded location (in the jumper task), or random location in the running (randRun) or jumper (randJumper) task. (C) Same as (B) but for the Jedi task. For Jedi, only periods when decoded location was near (within 5 cm of) the goal were included [also for (E)]. [(D) and (E)] Correlation of PV with rPV relative to baseline random value as a function of the time integration window for determining the PV. (F and G) Evaluation of decoder performance when ground truth activity for each location, i.e., the rPV, was input into the decoder. (F) Schematic of evaluation procedure. (G) Comparison of our DNN decoder to Bayesian decoder for different levels of added noise, with example traces using a specific level of noise (top). (H) Distribution of decoded location (left) during Jedi task segment with no treadmill movement (right). Right, from top: summed activity across all units with PBEs identified, treadmill speed, distance of decoded location (excluding data during PBEs) from goal, and close-up of LFP with spectrogram. (I) Power spectral density of z-scored (for pooling across animals) LFP during the Jedi task for periods of treadmill movement and all long segments (≥8 s) without treadmill movement. See text and methods in the supplementary materials for details. Here and elsewhere, all confidence intervals (CIs) are 95% CIs. Lai et al., Science 382, 566–573 (2023) 3 November 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E with the rPVs associated with the decoded loca- tions versus random locations, consistent with the hypothesis. Furthermore, jumper PV-rPV correlations were comparable to running task PV-rPV correlations. In line with this, the ex- ample running (Fig. 2B) and jumper (Fig. 2D) trials, which happened to share similar trajec- tories, showed similar activity patterns across place units over time. PV-rPV correlation scores were, unlike in jumper, lower in Jedi than in the running task using 500-ms windows (Fig. 4C), consistent with noisier generation of non- local representations and/or lower firing rates (fig. S13B) in Jedi. However, with longer in- tegration windows (>500 ms) (Fig. 4, D and E), the PVs generated during Jedi matched the rPVs as well as the best match during the running task (note that longer integration times work for Jedi because animals activated goal location representations for extended periods). These results indicate that, during BMI task performance, animals generated nonlocal pop- ulation activity as similar to the corresponding place field representations as when they ac- tually visited those locations in the running task. Were these place field–like (i.e., rPV-like) patterns what our deep network detected to decode location? While determining what fea- tures a deep network uses for decoding is gen- erally not straightforward, inputting a single location’s rPV for a brief duration was suffi- cient to produce accurate location decoding (Fig. 4, F and G), consistent with the decoder being tuned to detect rPV-like activity. In ad- dition, unlike the commonly used Bayesian decoder (45), our decoder was highly robust to noise (Fig. 4G) by design because of the use of data augmentation during training. Lastly, we analyzed the local field potential (LFP) activity during BMI task performance (Fig. 4, H and I). When animals move, the ro- dent hippocampal LFP is known to display prominent theta band (∼5 to 12 Hz) power, which peaked at ~7.3 Hz during periods of movement in the running and BMI tasks (Fig. 4I). During the extended periods of nonmove- ment when the animal was performing the Jedi task, the theta peak shifted down to 6.3 Hz (Fig. 4I). Note that, unlike the more continuous theta oscillations during movement, the oscil- lations during such nonmovement periods tended to be more intermittent. Discussion Previous BMI research has yielded major ad- vances in the control of robotic arms, com- puter cursors, and other devices by activity from the primary motor cortex, premotor cor- tex, and posterior parietal cortex (24–35). The hippocampal cognitive map has a code that represents space in terms of absolute location in the external environment versus location relative to (e.g., in front of, or to the right or left of) the animal (8–11), and it was unknown whether a subject could control a BMI by means of this code. With this study, we dem- onstrated a hippocampal map–based BMI in which the subject is able to control its location or that of other objects by activating location representations in terms of absolute space, in- dependent of where the animal currently is. That is, even though animals generally (but not always) turned their body toward the goal, the activity that needed to be generated dif- fered depending on the location of the goal with respect to the environment. The relative- ly small amount of training needed for the animals to perform our BMI tasks is in line with our use of a biomimetic decoder (35, 46), that is, one based on the neural code that the subject naturally employs. In humans, imagining or recalling objects or video clips is accompanied by hippocampal activity in individual neurons similar to that when viewing the original stimuli (47, 48). This suggests that the mechanisms allowing animals to selectively activate their nonlocal hippocampal spatial representations, as we have shown here, could also underlie our abil- ity to actively recall or imagine experiences in other places. The ability of rodents to perform these BMI tasks should thus allow imagina- tion, as well as the voluntary recall of memory, to be investigated using the range of tools available for this model system. More gener- ally, the neural processes engaged here could underlie our capacity to perform “mental time travel”—travel back in time by reexperiencing richly detailed episodic memories and travel forward in time by generating possible future scenarios (49). Mental time travel depends critically on the hippocampus (4–6, 50–52) and enables subjects to internally simulate new experiences according to their world model. This can aid decision-making and facilitate learning in complex situations where trial and error is expensive, as shown using artificial agents (3, 53–55). Along these lines, the rats in our study could control their hippocampal map-based activity on a timescale of seconds, corresponding to the speed and duration at which humans re- live past events or imagine new scenarios. Navigational trajectories each lasted ~10 s, and a virtual object could be held at a remote location for several seconds. This contrasts with the previously described fast (~100 ms) sequences of nonlocal hippocampal activity in awake rodents (i.e., awake replay events, which are associated with population bursts and sharp wave–ripples) thought to be associated with planning (12, 16, 18, 21), and which were not responsible for the performance in our BMI tasks (analysis in which all PBEs were re- moved). The content of such replay events, which can portray specific routes through the environment starting from the animal’s current location, has been shown to be corre- lated with deliberative (12) and future (16, 18, 21) behavior. However, it is not known whether this content is—or replay content in general can be—under an animal’s volitional control. For instance, hippocampal activity displays similar fast sequences during sleep (23), thus nonlocal path generation per se does not ap- pear to require intention. If awake replay is volitionally controlled, these events could rep- resent a brief consideration of alternatives for making a quick decision and be distinct from the more comprehensive mental simulations of possible scenarios that take seconds. Previ- ous work has also described neurons in the hippocampus and related areas whose activity is tuned to the angle to a goal or a salient cue or object relative to the direction the animal is facing (56–59). In addition, hippocampal neu- rons that are tuned to the location of con- specifics have been found (60, 61). As with fast sequences, whether these forms of activity that reflect locations away from the animal are volitionally controlled is yet to be determined. Beyond aiding decision-making, the ability to control the content of the hippocampal spatial and episodic memory system could help explain the richness of our inner lives. Finally, the ability to control hippocampal activity to guide oneself or objects to intended locations— and do so with high signal-to-noise readout using our decoder—could lead to new BMI ap- plications for restoring or enhancing function by realizing a subject’s high-level intentions with respect to their internal world models. REFERENCES AND NOTES 1. D. L. Schacter, D. R. Addis, R. L. Buckner, Nat. Rev. Neurosci. 8, 657–661 (2007). 2. D. L. Schacter et al., Neuron 76, 677–694 (2012). 3. 4. D. Hassabis, D. Kumaran, S. D. Vann, E. A. Maguire, Proc. Natl. J. B. Hamrick, Curr. Opin. Behav. Sci. 29, 8–16 (2019). Acad. Sci. U.S.A. 104, 1726–1731 (2007). 5. D. Hassabis, D. Kumaran, E. A. Maguire, J. Neurosci. 27, 14365–14374 (2007). 6. D. R. Addis, A. T. Wong, D. L. Schacter, Neuropsychologia 45, 1363–1377 (2007). 7. E. C. Tolman, Psychol. Rev. 55, 189–208 (1948). 8. J. O’Keefe, L. Nadel, The Hippocampus as a Cognitive Map (Clarendon Press, 1978). J. O’Keefe, J. Dostrovsky, Brain Res. 34, 171–175 (1971). 9. 10. J. O’Keefe, Prog. Neurobiol. 13, 419–439 (1979). 11. M. A. Wilson, B. L. McNaughton, Science 261, 1055–1058 (1993). 12. A. Johnson, A. D. Redish, J. Neurosci. 27, 12176–12189 (2007). 13. K. Diba, G. Buzsáki, Nat. Neurosci. 10, 1241–1242 (2007). 14. T. J. Davidson, F. Kloosterman, M. A. Wilson, Neuron 63, 497–507 (2009). 15. M. P. Karlsson, L. M. Frank, Nat. Neurosci. 12, 913–918 (2009). 16. B. E. Pfeiffer, D. J. Foster, Nature 497, 74–79 (2013). 17. A. M. Wikenheiser, A. D. Redish, Nat. Neurosci. 18, 289–294 (2015). 18. C.-T. Wu, D. Haggerty, C. Kemere, D. Ji, Nat. Neurosci. 20, 571–580 (2017). 19. K. Kay et al., Cell 180, 552–567.e25 (2020). 20. X. Mou, A. Pokhrel, P. Suresh, D. Ji, Neuron 110, 891–902.e7 (2022). 21. J. Widloski, D. J. Foster, Neuron 110, 1547–1558.e8 (2022). 22. K. Louie, M. A. Wilson, Neuron 29, 145–156 (2001). 23. A. K. Lee, M. A. Wilson, Neuron 36, 1183–1194 (2002). 24. J. Wessberg et al., Nature 408, 361–365 (2000). 25. J. M. Carmena et al., PLOS Biol. 1, E42 (2003). 26. S. Musallam, B. D. Corneil, B. Greger, H. Scherberger, R. A. Andersen, Science 305, 258–262 (2004). 27. L. R. Hochberg et al., Nature 442, 164–171 (2006). Lai et al., Science 382, 566–573 (2023) 3 November 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E 28. G. Santhanam, S. I. Ryu, B. M. Yu, A. Afshar, K. V. Shenoy, Nature 442, 195–198 (2006). 29. M. Velliste, S. Perel, M. C. Spalding, A. S. Whitford, A. B. Schwartz, Nature 453, 1098–1101 (2008). 30. P. T. Sadtler et al., Nature 512, 423–426 (2014). 31. T. Aflalo et al., Science 348, 906–910 (2015). 32. G. K. Anumanchipalli, J. Chartier, E. F. Chang, Nature 568, 493–498 (2019). 33. F. R. Willett, D. T. Avansino, L. R. Hochberg, J. M. Henderson, K. V. Shenoy, Nature 593, 249–254 (2021). 34. E. E. Fetz, J. Physiol. 579, 571–579 (2007). 35. M. A. Lebedev, M. A. L. Nicolelis, Physiol. Rev. 97, 767–837 (2017). 36. D. Ishikawa, N. Matsumoto, T. Sakaguchi, N. Matsuki, Y. Ikegaya, J. Neurosci. 34, 5044–5053 (2014). 37. K. Patel, C. N. Katz, S. K. Kalia, M. R. Popovic, T. A. Valiante, Brain 144, 3651–3663 (2021). 38. M. Cerf et al., Nature 467, 1104–1108 (2010). 39. D. Aronov, D. W. Tank, Neuron 84, 442–456 (2014). 40. C. Lai, thesis, University of Cambridge (2020). 41. C. Lai et al., bioRxiv 2023.09.14.557854 [Preprint] (2023); https://doi.org/10.1101/2023.09.14.557854. 42. C. D. Harvey, F. Collman, D. A. Dombeck, D. W. Tank, Nature 461, 941–946 (2009). 43. G. Chen, J. A. King, Y. Lu, F. Cacucci, N. Burgess, eLife 7, e34789 (2018). 44. J. S. Lee, J. J. Briguglio, J. D. Cohen, S. Romani, A. K. Lee, Cell 183, 620–635.e22 (2020). 45. K. Zhang, I. Ginzburg, B. L. McNaughton, T. J. Sejnowski, J. Neurophysiol. 79, 1017–1044 (1998). 46. K. V. Shenoy, J. M. Carmena, Neuron 84, 665–680 (2014). 47. G. Kreiman, C. Koch, I. Fried, Nature 408, 357–361 (2000). 48. H. Gelbard-Sagiv, R. Mukamel, M. Harel, R. Malach, I. Fried, Science 322, 96–101 (2008). 49. E. Tulving, Psychologie canadienne 26, 1–12 (1985). 50. F. Andelman, D. Hoofien, I. Goldberg, O. Aizenstein, M. Y. Neufeld, Neurocase 16, 426–435 (2010). 51. E. Race, M. M. Keane, M. Verfaellie, J. Neurosci. 31, 10262–10269 (2011). 52. T. Bartsch, J. Döhring, A. Rohr, O. Jansen, G. Deuschl, Proc. Natl. Acad. Sci. U.S.A. 108, 17562–17567 (2011). 53. T. Weber et al., Adv. Neural Inf. Process. Syst. 30, 5690–5701 (2017). 54. D. Ha, J. Schmidhuber, arXiv:1803.10122 [cs.LG] (2018). 55. D. Hafner, T. Lillicrap, J. Ba, M. Norouzi, arXiv:1912.01603 [cs.LG] (2019). 56. L. Acharya, Z. M. Aghajan, C. Vuong, J. J. Moore, M. R. Mehta, Cell 164, 197–207 (2016). 57. A. Sarel, A. Finkelstein, L. Las, N. Ulanovsky, Science 355, 176–180 (2017). 58. C. Wang et al., Science 362, 945–949 (2018). 59. J. Ormond, J. O’Keefe, Nature 607, 741–746 (2022). 60. T. Danjo, T. Toyoizumi, S. Fujisawa, Science 359, 213–218 (2018). 61. D. B. Omer, S. R. Maimon, L. Las, N. Ulanovsky, Science 359, 218–224 (2018). 62. C. Lai, S. Tanaka, T. D. Harris, A. K. Lee, Volitional activation of remote place representations with a hippocampal brain‐ machine interface, version 1.0, Zenodo (2023); https://doi.org/ 10.5281/zenodo.8360872. ACKN OWLED GMEN TS We thank M. Bolstad for work on the virtual reality software; W. Bishop, S. Romani, and X. Zhao for valuable discussions regarding the study; S. Lindo, R. Gattoni, and other members of the Janelia Vivarium team, and B. Lustig, A. Sohn, S. Sawtelle, I. Negrashov, B. Foster, J. Osborne, J. Arnold, R. Rogers, and J. Cohen for technical assistance; and J. Dudman, S. Romani, A. Hantman, T. Wang, and J. Colonell for valuable comments on the manuscript. Funding: This work was supported by the Howard Hughes Medical Institute (A.K.L. and T.D.H.). Author contributions: Conceptualization: C.L. and A.K.L. Task design: C.L. BMI and task software and BMI hardware: C.L. Behavioral and recording hardware: S.T. Experiments: S.T. and C.L. Data analysis: C.L. and A.K.L. Writing: C.L., A.K.L., and S.T. Supervision: A.K.L. and T.D.H. Competing interests: T.D.H. is also affiliated with the Department of Biomedical Engineering at Johns Hopkins University. Data and materials availability: All data needed to assess the findings in this study are publicly available at Zenodo (62). All analysis methods are described in the main text and supplementary materials, and the code for building and applying the decoder is publicly available at Zenodo (62). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse. This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the author-accepted manuscript (AAM) of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh5206 Materials and Methods Figs. S1 to S14 Table S1 References (63–72) MDAR Reproducibility Checklist Movie S1 Submitted 9 March 2023; accepted 22 September 2023 10.1126/science.adh5206 Lai et al., Science 382, 566–573 (2023) 3 November 2023 8 of 8
10.1126_science.adh5315
RES EARCH QUANTUM OPTICS Quantum vortices of strongly interacting photons Lee Drori†, Bankim Chandra Das†, Tomer Danino Zohar, Gal Winer, Eilon Poem, Alexander Poddubny, Ofer Firstenberg* Vortices are topologically nontrivial defects that generally originate from nonlinear field dynamics. All-optical generation of photonic vortices—phase singularities of the electromagnetic field—requires sufficiently strong nonlinearity that is typically achieved in the classical optics regime. We report on the realization of quantum vortices of photons that result from a strong photon-photon interaction in a quantum nonlinear optical medium. The interaction causes faster phase accumulation for copropagating photons, producing a quantum vortex-antivortex pair within the two-photon wave function. For three photons, the formation of vortex lines and a central vortex ring confirms the existence of a genuine three-photon interaction. The wave function topology, governed by two- and three-photon bound states, imposes a conditional phase shift of p per photon, a potential resource for deterministic quantum logic operations. P hotons do not interact with one another in the optical regime. An effective inter- action between photons can be medi- ated by matter, but, in conventional optical nonlinear media, this interac- tion is negligible on the level of individual photons. It is only at the ultimate limit of quan- tum nonlinear optics, realized in specially en- gineered systems, that a single photon can alter the optical response of the system, ren- dering an appreciable photon-photon inter- action (1–4). The realization of strong photon-photon in- teractions in atomic ensembles can enable de- terministic quantum-information processing for optical qubits, such as single-photon tran- sistors and phase gates (5–7). It can be used to generate nonclassical states of light, such as squeezed, cluster, and repeater states, as a resource for quantum communication, com- putation, and sensing (8, 9). It can also mani- fest in interacting quantum gases and fluids of photons, allowing the exploration of exotic many-body physics for light (10–13). We realize an extreme regime of quantum nonlinear optics and observe optical quantum vortices generated by the interaction between only two or three photons. Quantum vortices— phase singularities of the wave function—are typically expected in strongly interacting sys- tems of many particles, such as nonlinear op- tical systems (14, 15), superfluids (16–19), and hybrid light-matter systems (20–24). In this work, we induce a strong, effective interaction at the few-photon level by coupling photons to Rydberg atoms in an ultracold rubidium gas (Fig. 1, A and B). The photons propagate through the me- dium as light-matter polaritons, at 10−6 times the speed of light. Because of the strong van der Waals Department of Physics of Complex Systems, Weizmann Institute of Science, Rehovot 7610001, Israel. †These authors contributed equally to this work. *Corresponding author. Email: ofer.firstenberg@weizmann.ac.il Drori et al., Science 381, 193–198 (2023) 14 July 2023 coupling between the Rydberg atoms, photons that propagate close to each other—at a dis- tance smaller than the so-called blockade radius rb—experience a local change in the refractive index. For two photons, the optical field can be described by a wave function y(x1, x2), which is the probability amplitude of having photons at coordinates x1 and x2 along the medium (25). The change in refractive index at x1 (cid:1) x2 j ≤ rb causes the accumulation of excess local phase in y(x1, x2), which, under the correct conditions, produces vortices in y(x1, x2). j To understand the vortex formation, we con- sider a simplified analytical model describing the evolution of y(x1, x2) in a medium of uni- form density and length L using the two-photon Schrödinger equation (25, 26): i@Ry ¼ (cid:1) 1 2m @2 r y þ V 2ð Þ rð Þy ð1Þ Here, r = x1 − x2 is the relative coordinate of the two photons, and the center-of-mass coordi- nate along the medium R = (x1 + x2 + L)/2 = 0… L plays the role of time. The incoming, co- herent probe field dictates homogeneous initial conditions y(r, R = 0) = 1. The photon-photon interaction is described by a nearly square 6), where U con- potential V (2)(r) = U/(1 + r6/rb stitutes the change in the refraction index due to the Rydberg blockade, and the (negative) effective mass m = −U/8 arises from single- photon dispersion. Because mU < 0, Eq. 1 de- scribes attraction between photons (25). A nontrivial prediction of Eq. 1 is the for- mation of vortex-antivortex pairs in the me- dium, symmetrically around the potential well (Fig. 1C). Equation 1 has localized and extended eigenstates, denoted as bound and scattering two-photon states (25). Considering a single bound state ybound(r) with energy E2, which is the case in our experiment, the initial scatter- ing component is yscat = 1 – ybound(r). In the crudest approximation, the phase accumula- tion of the scattering component can be ne- glected (27), and the solution to Eq. 1 reduces to y r; Rð Þ ¼ e(cid:1)iE2Ry bound rð Þ þ y scat ð2Þ The phase vortices are formed around y = 0 when the two terms—the bound and scatter- ing components—cancel each other. This re- sult is generic and occurs for a sufficiently long medium and for any number of bound states. Such an interference mechanism for vortex- antivortex pair formation is universal and known from acoustics (28) and atomic col- lisions (29) to linear (30, 31) and nonlinear (32) optics. However, such vortex-antivortex pair formation has not been observed before in quantum optics because the required strong and prolonged photon-photon interaction had been unattainable. jr2 We quantify the duration and strength of the interaction by the dimensionless parame- ters ϕ = UL/2 and l ¼ 2 Umj b (33), respec- tively. For vortices to develop, ϕl should be large. When the interaction is weak, that is, for l ≪ 1, we find the condition ϕl > ϕ0, with the threshold phase ϕ0 ≈ 0.94p. For moderate in- teractions l ≥ 1, the condition is ϕ > ϕ0 [see section S2.4 in (27)]. These conditions are illustrated by solid lines in the “phase diagram” in Fig. 1D, and they compare reasonably well with the domain of vortex formation calculated numerically for our experimental, nonuniform medium (green shading). In Rydberg polariton systems, the essen- tial experimental parameters are the total optical depth OD of the medium (e−OD being the on-resonance transmission) and the optical depth of the blockade range ODb = OD·rb/L. These parameters govern ϕ º OD and l º 2. In our experiment, we reach OD ≈ 110 ODb and ODb ≈ 22, setting ϕ ≈ 2.7p and l ≈ 3 (blue diamond in Fig. 1D). The moderate inter- action obtained in previous experiments (ϕ ≈ 1.5p and l ≪ 1), albeit being enough to ob- serve the signatures of photonic bound states (25, 34, 35), were well below the threshold for vortex formation. Observation of two-photon vortices p Our experiment starts by compressing and trapping an ultracold rubidium cloud with a 0e(cid:1)x2=2s2 centered at Gaussian density profile r ffiffiffiffiffi x = 0 and effective length L ¼ s ≈ 75 mm 2p along the optical axis x. Our maximal peak density r0 ≈ 5 × 1012 atoms/cm3 is a factor of three to six higher than that in previous ex- periments and is the primary source of our large l º r0 2. We form Rydberg polaritons using counter-propagating probe and con- trol fields, which together resonantly couple the atomic 5S ground level to the 100S Rydberg level via the 5P level (Fig. 1B). We send, on av- erage, f ≈ 0.25 probe photons per microsecond. The probe photons experience the three- level optical susceptibility, except within the 1 of 6 RES EARCH | R E S E A R C H A R T I C L E A C B D Fig. 1. Setup and conditions for generating photon vortices in Rydberg- mediated quantum nonlinear optics. (A) Counter-propagating probe (red) and control (blue) fields are focused onto an elongated ultracold atomic cloud (left, absorption image), held in a crossed optical trap (green). After traversing the medium, the probe light is split and measured by four single-photon detectors D1 to D3 and Df. Detectors D1 to D3 provide the two-photon and three-photon correlation functions g(2) and g(3). Detector Df, which measures the interference beat between the probe and reference light (LO, local oscillator), is correlated with detectors D1 to D3 to extract the conditional phases f(2) and f(3) (see also fig. S2). (B) Atomic-level structure for generating polaritons comprising the 100S Rydberg orbital. The van der Waals interaction (vdW) between two Rydberg atoms disturbs the propagation of two close polaritons, leading to local p ffiffiffiffiffi 12 accumulation of excess propagation phase. (C) Amplitude and phase of the two-photon wave function, calculated numerically in the Schrödinger approxi- mation (Eq. 1, with U ¼ =rb), showing the formation of a vortex-antivortex pair of two interacting photons. (D) Conditions for vortex generation (green shaded area, calculated numerically for a realistic Gaussian cloud) in terms of the photon-photon interaction strength l and the accumulated interaction phase in the finite medium ϕ. The solid curves are minimal conditions calculated analytically for a uniform cloud under the approximation of Schrödinger evolution (27). The dotted blue line indicates the available ϕ and l in our experimental setup, peaking at the blue diamond. Red circles indicate the conditions achieved in (25, 34, 35). The inset image depicts the phase structure of a single vortex-antivortex pair. blockade range rb ≈ 15.3 mm, where they experience the two-level (5S-5P) susceptibility. Ideally, we want the difference U between these susceptibilities to be purely real, render- ing a conservative photon-photon interaction. To this end, we first detune the probe from the atomic transition by D = 9.4G, where G is the half-width of the 5P level, and then adjust the control frequency such that the transmis- sion remains the same outside and inside of the blockade range (25). This adjustment elimi- nates the residual nonconserving (dissipative) part of the interaction and amplifies the con- serving part U = (OD/L)(qG/D) by the factor q ≈ 1.4 (27). The atomic density decreases dur- ing the experimental cycle by a factor of 5.5, allowing us to study the full range 110 ≥ OD ≥ 20 (2.7p ≥ ϕ ≥ 0.5p) and, correspondingly, 22 ≥ ODb ≥ 4.1 (3 ≥ l ≥ 0.1) in each experiment (dotted line in Fig. 1D). The vortex-antivortex pair forms in the two- photon wave function y(r, R) of the probe inside the medium. Although we cannot di- rectly measure these vortices within the medium, we experimentally determine their presence and structure from the correlations of the outgoing photons, essentially measuring y(r, R = L) at the medium’s boundary. We track the depen- dence of y(r, R = L) on the optical depth OD, and by varying OD, we effectively vary the interaction duration. The dependence on OD at the boundary becomes a proxy for the evo- lution in the bulk. Whereas OD corresponds to R, the temporal separation t between the outgoing photons coarsely corresponds to their spatial separation r in the medium (36). Ex- perimentally, we measure the two-photon cor- relation function g(2)(t) and conditional phase f(2)(t) of the outgoing probe field for varying OD and associate them, respectively, with the squared amplitude and phase of y(r, R = L). In the experimental results (Fig. 2, A and B), we first observe the gradual bunching of pho- tons, g(2)(0) > 1, as the OD increases, accom- panied by depletion of g(2)(t) at tj j ≈ 0:25 ms [see cross section (i) in Fig. 2A]. The bunching and depletion are due to the effective attraction between the photons, which is governed by a two-photon bound state and accompanied by accumulation of conditional phase f(2)(0) < 0 [see (i) in Fig. 2B] (25). At OD ≈ 78, f(2)(0) reaches −p. A p conditional phase allows for deterministic, maximally entangling operation on photonic qubits (7), and here it is obtained for copropagating photons. The vortex-antivortex pair is formed around OD ≈ 80, as manifested by the clockwise and counterclockwise phase twists in Fig. 2B (or- ange arrows). These phase twists correspond to steep steps of f(2)(t) at tj j ≈ 0:25 ms, whose direction flips when the OD increases [com- pare (ii) with (iii) in Fig. 2B]. Furthermore, as expected, the photons are depleted from the vortex core, where g(2)(t) = 0.24 (global minima in Fig. 2A). Following the vortex-antivortex for- mation, f(2)(0) decreases gradually from p to zero, and the bunching and depletion in g(2) get smaller. The vortices effectively “unwind the tension” between the regions of fast and slow accumulation of phase. Our experimental results are also corrobo- rated by numerical simulations based on the model described in (25, 36). The simulations use the actual experimental parameters, in- cluding the Gaussian density profile of the Drori et al., Science 381, 193–198 (2023) 14 July 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E A B C Fig. 2. Two-photon vortices. (A and B) Color maps show the experimental (A) two- photon correlation g(2) and (B) two-photon phase f(2) as a function of the time between photons t and the optical depth of the atomic medium OD. Enlarged views (i) to (iii) explicitly show the measured curves g(2)(t) and f(2)(t) for OD = 49, 78, and 95. A vortex-antivortex pair [orange arrows in (B)] is formed around OD = 80 on the edge of the medium and is thus captured by the detectors as a phase step around |t| = 0.25 ms and a phase wrap of f(2)(t = 0) from −p to +p. At the vortices’ core, g(2) approaches zero. (C) Numerically calculated f(2) for the experimental conditions. (D and E) Nu- merically calculated phase of the stationary, space-dependent two- photon wave function, showing the two-photon vortices forming (D) deep inside the medium and (E) on the edge. Dashed circles represent the edge (2s) of the Gaussian cloud. D medium and the propagation of the second photon after the detection of the first photon. They are presented in Fig. 2C and are in excellent agreement with the measured data, reproduc- ing the vortex-antivortex pair and the surround- ing phase structure in full detail. The 10% discrepancy in the OD at which the vortices are observed is due to a slight saturation of our phase detection scheme and is completely resolved for a weaker incoming probe (see fig. S5). We use these simulations to calculate the conditions for vortex formation in Fig. 1D (green shaded area). Moreover, the simulations are able not only to reproduce the experimentally accessible data but also to establish the dynamics of the two- photon wave function y(x1, x2) inside the me- dium. Figure 2, E and D, presents the phase of y(x1, x2) for two values of OD. These simu- lations exhibit the entrance of phase singulari- ties and eventually the presence of a stationary pair of quantum vortices in the medium. Three-photon vortices The strong interaction, responsible for forming two-photon vortex-antivortex pairs, leads to an even richer topological structure of the three- photon wave function y(x1, x2, x3). We will first illustrate this by considering only the pairwise photon interactions. If only two of the three photons are close together, they attract each other and form a quasi-bound two-photon state. Because the third photon remains unbound, the three pairs of vortices resulting from each of the three possible quasi-bound states would mani- fest as six vortex lines in the (x1, x2, x3) space. To verify this prediction, we measured the three-photon phase f(3)(t1, t2, t3) and the cor- responding correlation function g(3)(t1, t2, t3) over a range of OD. Whereas the stationary two-photon correlation function depends on a single time separation t1 − t2, characteriza- tion of the three-photon wave function requires two time separations. These are conveniently parametrized by the Jacobi coordinates h ¼ t21= ð and z ¼ t13 þ t23 , where tij = ti − tj (37). The three-photon measure- ments are then given by three-dimensional (3D) datasets for f(3) and g(3) that depend on h, z, and OD. ffiffiffi 6 ffiffiffi 2 Þ= p p (cid:3) (cid:3) Figure 3A shows f(3)(h, z) at the critical OD where vortices appear. We identify cores of vortex lines with p-phase dislocations: along (cid:3) (cid:3) ¼ p region) the edges of the six-bar star ( f 3ð Þ and around the center (f(3) = 0 region). Figure 3, B and C, shows the development of the phase along the lines z = 0 (uniformly separated photons) and h = 0 (a photon at some distance from a pair). Notably, f(3)(0, 0) varies mono- tonically with OD much faster than f(2)(0) does, as seen by comparing Fig. 3B with Fig. 2B. Eventually, the vortices structure is best understood from Fig. 3D, which shows the re- (cid:3) (cid:3) ¼ 0:7 rad=ms, constructed 3D isosurfaces ∇f 3ð Þ with the phase f(3) indicated by the surface color. The starlike structure of tubes in Fig. 3D corresponds to six vortex lines, in agree- ment with our prediction. We can addition- ally verify that this structure is due only to the two-photon attraction by constructing it from the measured two-photon data. This construc- tion, shown in fig. S4, exhibits a starlike vortex structure nearly identical to that found in the three-photon data. (cid:3) (cid:3) ¼ y Þ þ y Þ þ y bound r21ð bound r32ð bound r13ð The vortex star structure can be explained analytically by generalizing Eq. 2 to three pho- tons. To this end, we replace ybound(r) by a sum y 2→3 ð Þ Þ of bound three quasi-bound states (rij = xi – xj), respecting the bosonic (permutation) symmetry of the three- photon wave function. For simplicity, we assume bound rð Þ ¼ 2e(cid:1)2 rj j=a in a a weakly bound state y delta potential, where a ≈ rb/l is the scattering length (25, 38). As shown in Fig. 3E using the spatial Jacobi coordinates h ¼ r21= and ffiffiffi z ¼ r13 þ r23 , the generalized Eq. 2 gives 6 the six vortex lines. These results support our ffiffiffi 2 Þ= p p ð Drori et al., Science 381, 193–198 (2023) 14 July 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Three-photon vortex lines and vortex ring. (A to D) Three-photon con- ditional phase f(3)(h, z) measured at different OD, where h and z are time separations between the photons in Jacobi coordinates. For each OD, the data are averaged by assuming a sixfold symmetry; an example of data before averaging is shown in the inset in (A). We present three cross sections: (A) OD = 79, (B) z = 0, and (C) h = 0. In all these, the central feature originates from three-photon interactions. In (D), to reveal the vortex structure, we plot the measured f(3) along (cid:3) (cid:3) (cid:3) (cid:3), which are themselves derived from the data. We identify six isosurfaces of rfð3Þ vortex lines (tubes) and a central vortex ring (torus); note the 2p phase twist around the cross section of the tubes and torus (inset). (E and F) Analytic calculation of the three-photon phase arg[y(h, z, R)] using the approximate ansatz in Eq. 3 (E) without and (F) with the three-photon bound-state term, for E3 = 3E2 and y 3ð Þ Þ. The outer vortex lines are due to the generalized two- bound 0; 0ð photon bound states, whereas the vortex ring is due to the three-photon bound state. Þ ¼ (cid:1)3y 2→3 bound 0; 0ð Þ ð interpretation that the vortex lines in the three-photon wave function are a direct gen- eralization of the two-photon vortices. The full three-photon data, however, show richer behavior, which cannot be reduced to just independent pairwise photon attraction. The effect of a third photon approaching a bound pair is reflected in the central feature in Fig. 3C and, more clearly, as a torus around h = z = 0 in Fig. 3D, manifesting a vortex ring. This inner ring does not originate from two-photon quasi-bound states but rather from a genuine three-photon bound state y 3ð Þ bound (34, 37) that interferes with the scattering states. To describe the formation of the ring, we bound and its use the following ansatz, adding y 3ð Þ energy E3 to the generalized Eq. 2: y h; z; R ð Þ ¼ e(cid:1)iE3Ry 3ð Þ h; z ð bound ð Þ þ e(cid:1)iE2Ry 2→3 bound Þ þ y scat ð3Þ Þ ð scat bound bound h; z p Þº e(cid:1) ¼ 1 (cid:1) y 3ð Þ ffiffi ffiffi p p j hþ j=ae(cid:1) z 2 3 ð (cid:1) y 2→3 wherey bound . As shown in (34, 39), the wave function of a three-photon weakly bound state near h = z = 0 can be approximated ffiffi ffiffi p byy 3ð Þ j=a. j h(cid:1) jhj=ae(cid:1) z 2 8 This form can be interpreted as a product of confined states, with three pairs attracting each other simultaneously. It is then straight- forward to check, as shown in Fig. 3F, that the interference between the first and last terms in Eq. 3 produces a vortex ring with the same topology that we observe experimentally. ffiffi p 3 Another important result revealed by the f(3) data in Fig. 3 is that the vortex lines and ring appear approximately at the same value of OD ≈ 80. This observation is far from obvious, allowing us to estimate the relation between the bound state energies E3 and E2. An analytical model with only pairwise in- teraction V (2) predicts E3 = 4E2 (40), which would result in the vortex ring appearing at j a lower OD. Fitting Eq. 3 to the experimental data yields E3 = 3E2, so that e(cid:1)iE3R ¼ (cid:1)1 forms a vortex ring simultaneously with e(cid:1)iE2R ¼ (cid:1)1 forming the vortex lines, as shown in Fig. 3F. The result E3=E2 j < 4 is experimental evi- dence of a genuine three-photon repulsion term V (3)(h, z) that attenuates the pairwise photon attraction (37, 41). This attenuation follows from the physics of the Rydberg block- ade, where one photon can “saturate” the in- teraction and simultaneously block two (or more) other photons. The conditional (same-time) phase f 3ð Þ 0 ≡ f 3ð Þ 0; 0ð Þ, which originates from the construc- tive sum of the vortex lines and ring, is f 3ð Þ ≈ 0 (cid:1)2p. This nonzero value is evident from the monotonic increase of the phase versus OD in Fig. 3, B and C, or from the phase winding directions in Fig. 3D. Alternatively, one can unwrap the f(3) data, as shown in fig. S3, to find −2p at the center of the ring. Impor- tantly, given that f 2ð Þ ≡ f 2ð Þ 0ð Þ≈ (cid:1)p, we find 0 f 3ð Þ 0 . This result deviates from the 0 known nonlinear quantum-Kerr phase for n photons f nð Þ Þ=2, which ascribes ð Kerrn n (cid:1) 1 Kerr f 3ð Þ ð2Þ ¼ (cid:1)p (42). Our result Kerr Kerr ð3Þ ¼ 2f 2ð Þ f 0 agrees with the prediction for sat- 0 urated interaction (34), again indicating the attenuation of the interaction by the block- ade saturation. ¼ f 2ð Þ ¼ (cid:1)3p for f ¼ 2f 2ð Þ The strong three-photon attraction in the l > 1 regime and the repulsive effect of the blockade saturation are also evident in the three-photon correlations g(3) presented in Fig. 4. Whereas the six crests of g(3) > 1 away from the center are governed by the two-photon quasi-bound states, the structure at the center is substantially modified by the three-photon bound state. To see this, compare the mea- sured g(3) data in Fig. 4, A to C, with that expected from only the pairwise interactions Þ þ g 2ð Þ t13ð (3) = g (3) − gd g 3ð Þ Þ þ g 2ð Þ t32ð ¼ g 2ð Þ t21ð Þ (cid:1) 2 in Fig. d 4, D to F. g 3ð Þ d is the so-called disconnected part of the correlations (37). It is evident in all cross sections that the central peak in g(3), reaching a maximal value of g(3)(0, 0) = 7.4 ± 0.3, is twice as narrow as the corresponding peak in g 3ð Þ d . This strong three-photon bunching has been characterized before and attributed to the tighter confinement of y 3ð Þ bound compared with y 2ð Þ bound (34, 37). The effect of blockade saturation is best captured by the connected part of the correla- (3) in Fig. 4, H and I. We tions gc (3) as a manifestation of the can interpret gc contribution of the three-photon correction term V (3), which is repulsive. First, the repul- sion attenuates the bunching of simultaneous (3) close photons, as evident by the negative gc to the center (blue in Fig. 4, G to I). We observe a minimal value ofg 3ð Þ ¼ (cid:1)2:5 T 0:2. Second, as photons are less attracted to the center, we (3) (at a radius observe a ring of positive gc p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi h2 þ z2 ≈ 0:25 ms; red in Fig. 4G), reaching g 3ð Þ c;max ¼ 0:9 T 0:1. As expected, with l > 1, both c;min and g 3ð Þ g 3ð Þ c;max are an order of magnitude larger than the values obtained previously with l ≪ 1 (34, 35, 43). Finally, gc (3) is modulated along that ring, exhibiting six peaks where the three photons are uniformly separated in time (e.g., along h, for z = 0). This, too, is due to the repulsive correction, which favors equally spaced photons over the asymmetric arrange- ment of one photon separated from a close pair (e.g., along z, for h = 0). For larger photon separations, the regularization effect of V (3) diminishes, and gc (3) approaches zero. c;min Discussion and summary Our observation of topological defects—including quantum vortex-antivortex pairs, vortex lines, and vortex rings—in the few-body wave function Drori et al., Science 381, 193–198 (2023) 14 July 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Three-photon bound-state and blockade saturation. (A to C) Three-photon correlations g(3)(h, z, OD) (after sixfold averaging), sectioned at (A) OD = 69, (B) z = 0, and (C) h = 0. (D to F) Disconnected part of the correlations gd calculated from only the g(2)(t, OD) data and representing the two-photon contribution to the three-photon correlations. The strong confinement of the three-photon (3)(h, z, OD), (3) = g(3) − gd bound state leads to a tighter bunching feature. (G to I) Connected part of the (3), reflecting the dynamics beyond pairwise interactions. correlations gc These dynamics are regulated by the saturation of the blockade interaction, expressed as an effective, short-range repulsive force. The negative central feature (blue), the positive ring (red), and the modulation along this ring are all governed by this force. of interacting photons is enabled predominantly by realizing a strong interaction between the photons. The interaction strength is quantified by the dimensionless parameter l, which is the ratio between the characteristic interaction and kinetic energies. Whereas in previous works this interaction strength was less than unity, our system reaches l ≈ 3. The vortices in the n-photon wave function (n) determine the maximal conditional phase f0 achievable for copropagating photons. For two photons, the vortex-antivortex pair develops (2) ≈ −p. For three photons, these pairs when f0 connect to form six vortex lines surrounding a (3) ≈ −2p. vortex ring, the sum of which gives f0 The deviation from the value −3p expected for a quantum-Kerr nonlinearity is a manifestation of the blockade saturation. It demonstrates the role played by the vortices in controlling the conditional phase, with important consequences for deterministic quantum logic operations (44, 45) and other quantum nonlinear devices (46, 47) with three or more photons. Our findings far from exhaust the potential of exploring many-body polariton physics. Even for relatively weak interactions, one could imagine exotic many-body bound states, such as a cascade of Efimov three-body states with a scaling-invariant spectrum. Although ini- tially proposed for 3D structures (48), Efimov states were later predicted to exist for a lower number of spatial dimensions (49, 50), including quasi–1D Rydberg atom clouds (13), but have not been observed thus far. Fur- ther, our observation of the genuine three- body interaction V (3) could be key to the realization of four-body interactions, which are important for lattice simulations of gauge field theories (51). As opposed to real space, which is limited to three dimensions, the (x1, x2, …) space studied in this work can be generalized to more than three dimensions, where higher-dimensional topological structures, such as linked and knotted vortex loops, could be explored. This could provide an interesting connection to the stud- ies of nontrivial topological physics in syn- thetic dimensions (52, 53). It could further be interesting to examine whether and how the topological phase singularities of the n-photon wave function transform to vortices of the mean field in the nonlinear classical optics description at n ≫ 1. Finally, a twofold increase in the atomic density will realize the regime l > p2, where higher-order two- and three-photon bound states should appear (26, 38). A superposition of vortex series arising from different bound states would drive complex nonperiodic dynam- ics of the few-photon wave function with an intricate phase structure. The three-photon vortex rings that we observe are a relatively simple example of how the phase of the n- photon wave function can be controlled by an additional (n + 1) photon. Potentially, these topological phase structures of n-photon wave function could be harnessed to develop new multiphoton control tools and deterministic quantum logic. Drori et al., Science 381, 193–198 (2023) 14 July 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E RE FE RENCES AND N OT ES 27. Materials and methods are available as supplementary 1. A. Imamog¯lu, H. Schmidt, G. Woods, M. Deutsch, Phys. Rev. Lett. 79, 1467–1470 (1997). 2. S. E. Harris, L. V. Hau, Phys. Rev. Lett. 82, 4611–4614 (1999). 3. D. E. Chang, V. Vuletić, M. D. Lukin, Nat. Photonics 8, 685–694 materials. 28. J. F. Nye, M. V. Berry, F. C. Frank, Proc. R. Soc. London Ser. A 336, 165–190 (1974). 29. J. O. Hirschfelder, K. T. Tang, J. Chem. Phys. 64, 760–785 52. A. Dutt et al., Science 367, 59–64 (2020). 53. A. Balčytis et al., Sci. Adv. 8, eabk0468 (2022). 54. L. Drori et al., Quantum vortices of strongly interacting photons. Zenodo (2023); https://doi.org/10.5281/zenodo. 7962196. (2014). 4. O. Firstenberg, C. S. Adams, S. Hofferberth, J. Phys. At. Mol. Opt. Phys. 49, 152003 (2016). 5. Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi, H. J. Kimble, Phys. Rev. Lett. 75, 4710–4713 (1995). 6. H. Gorniaczyk, C. Tresp, J. Schmidt, H. Fedder, S. Hofferberth, Phys. Rev. Lett. 113, 053601 (2014). 7. T. Stolz et al., Phys. Rev. X 12, 021035 (2022). 8. J. Hinney et al., Phys. Rev. Lett. 127, 123602 (2021). 9. C.-W. Yang et al., Nat. Photonics 16, 658–661 (2022). I. Carusotto, C. Ciuti, Rev. Mod. Phys. 85, 299–366 (2013). 10. 11. J. Otterbach, M. Moos, D. Muth, M. Fleischhauer, Phys. Rev. Lett. 111, 113001 (2013). 12. M. F. Maghrebi et al., Phys. Rev. A 91, 033838 (2015). 13. M. J. Gullans et al., Phys. Rev. Lett. 119, 233601 (2017). 14. V. Tikhonenko, J. Christou, B. Luther-Davies, Y. S. Kivshar, Opt. Lett. 21, 1129–1131 (1996). 15. A. V. Mamaev, M. Saffman, A. A. Zozulya, Phys. Rev. Lett. 76, 2262–2265 (1996). 16. G. Gauthier et al., Science 364, 1264–1267 (2019). 17. S. P. Johnstone et al., Science 364, 1267–1271 (2019). 18. J. R. Abo-Shaeer, C. Raman, J. M. Vogels, W. Ketterle, Science 292, 476–479 (2001). 19. L. Klaus et al., Nat. Phys. 18, 1453–1458 (2022). 20. J. Klaers, J. Schmitt, F. Vewinger, M. Weitz, Nature 468, 545–548 (2010). 21. E. Busley et al., Science 375, 1403–1406 (2022). 22. J. Kasprzak et al., Nature 443, 409–414 (2006). 23. K. G. Lagoudakis et al., Nat. Phys. 4, 706–710 (2008). 24. G. Roumpos et al., Nat. Phys. 7, 129–133 (2010). 25. O. Firstenberg et al., Nature 502, 71–75 (2013). 26. M. Moos, R. Unanyan, M. Fleischhauer, Phys. Rev. A 96, 023853 (2017). (1976). 30. K. Y. Bliokh, F. Nori, Phys. Rev. A 86, 033824 (2012). 31. K. Y. Bliokh, Phys. Rev. Lett. 126, 243601 (2021). 32. N. Jhajj et al., Phys. Rev. X 6, 031037 (2016). 33. P. Bienias, H. P. Büchler, New J. Phys. 18, 123026 (2016). 34. Q.-Y. Liang et al., Science 359, 783–786 (2018). 35. D. P. Ornelas-Huerta et al., Phys. Rev. Lett. 126, 173401 (2021). 36. T. Peyronel et al., Nature 488, 57–60 (2012). 37. K. Jachymski, P. Bienias, H. P. Büchler, Phys. Rev. Lett. 117, 053601 (2016). 38. P. Bienias et al., Phys. Rev. A 90, 053804 (2014). 39. Z. Chen, Y. Zhou, J.-T. Shen, Opt. Lett. 45, 2559–2562 (2020). 40. J. B. McGuire, J. Math. Phys. 5, 622–636 (1964). 41. M. J. Gullans et al., Phys. Rev. Lett. 117, 113601 (2016). 42. S. Haroche, J.-M. Raimond, Exploring the Quantum: Atoms, Cavities, and Photons (Oxford Univ. Press, 2006). 43. N. Stiesdal et al., Phys. Rev. Lett. 121, 103601 (2018). 44. X. Wang, A. Sørensen, K. Mølmer, Phys. Rev. Lett. 86, 3907–3910 (2001). 45. C.-P. Yang, Y. Liu, F. Nori, Phys. Rev. A 81, 062323 (2010). 46. M. R. Geller, Sci. Rep. 11, 402 (2021). 47. T. Opatrný et al., Sci. Adv. 9, eadf1070 (2023). 48. P. Naidon, S. Endo, Rep. Prog. Phys. 80, 056001 (2017). 49. Y. Nishida, S. Tan, Few-Body Syst. 51, 191–206 (2011). 50. S. Moroz, J. P. D’Incao, D. S. Petrov, Phys. Rev. Lett. 115, 180406 (2015). 51. R. Irmejs, M. C. Banuls, J. I. Cirac, Quantum simulation of Z2 lattice gauge theory with minimal resources. arXiv:2206.08909 [quant-ph] (2022). AC KNOWLED GME NTS We thank C. Avinadav for helpful discussions. Funding: We acknowledge financial support from the Israel Science Foundation, the US-Israel Binational Science Foundation (BSF) and the US National Science Foundation (NSF), European Research Council starting investigator grant QPHOTONICS 678674, the Minerva Foundation with funding from the Federal German Ministry for Education and Research, the Estate of Louise Yasgour, and the Laboratory in Memory of Leon and Blacky Broder. Author contributions: L.D., B.C.D., G.W., and O.F. conceptualized the project. L.D., B.C.D., T.D.Z., and G.W. performed the experiment and analyzed the data. B.C.D. and A.P. lead the theory work. All authors contributed to building the methodology and writing the final manuscript. B.C.D., E.P., A.P., and O.F. supervised the project. Competing interests: The authors have no competing interests. Data and materials availability: All data are available in the manuscript or the supplementary materials or are deposited at Zenodo (54). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh5315 Materials and Methods Supplementary Text Figs. S1 to S6 References Submitted 9 March 2023; accepted 7 June 2023 10.1126/science.adh5315 Drori et al., Science 381, 193–198 (2023) 14 July 2023 6 of 6
10.1126_science.adg8802
RES EARCH CELLULAR IMMUNOLOGY Perforin-2 is a pore-forming effector of endocytic escape in cross-presenting dendritic cells Pablo Rodríguez-Silvestre1, Marco Laub1, Patrycja A. Krawczyk1, Alexandra K. Davies2†, Julia P. Schessner2, Reejuana Parveen1, Benjamin J. Tuck1,3, William A. McEwan3, Georg H. H. Borner2, Patrycja Kozik1* During initiation of antiviral and antitumor T cell–mediated immune responses, dendritic cells (DCs) cross-present exogenous antigens on major histocompatibility complex (MHC) class I molecules. Cross-presentation relies on the unusual “leakiness” of endocytic compartments in DCs, whereby internalized proteins escape into the cytosol for proteasome-mediated generation of MHC I–binding peptides. Given that type 1 conventional DCs excel at cross-presentation, we searched for cell type– specific effectors of endocytic escape. We devised an assay suitable for genetic screening and identified a pore-forming protein, perforin-2 (Mpeg1), as a dedicated effector exclusive to cross-presenting cells. Perforin-2 was recruited to antigen-containing compartments, where it underwent maturation, releasing its pore-forming domain. Mpeg1−/− mice failed to efficiently prime CD8+ T cells to cell-associated antigens, revealing an important role for perforin-2 in cytosolic entry of antigens during cross-presentation. T he integrity of endosomal and lysosomal membranes is critical to protect the cell against extracellular pathogens and toxins, and from the activity of lysosomal hydro- lases. In dendritic cells (DCs), however, internalized proteins are delivered from endo- cytic organelles into the cytosol, where they can be proteolytically processed for presentation on major histocompatibility complex (MHC) class I molecules (1). The ability of DCs to present exogenous peptides on endogenous MHC-I is termed cross-presentation. Cross-presentation is critical for initiation of cytotoxic T cell (CTL) responses to antigens not expressed in DCs such as neoantigens or antigens from virally infected cells (2–4). Various mechanisms have been suggested to facilitate endocytic escape and promote cross-presentation (5–10). Early studies pro- posed that escape is mediated by protein channels (such as Sec61) recruited from the endoplasmic reticulum to endosomes (7, 11). More recent data suggest that escape occurs through unrepaired damage to membranes (e.g., due to reactive oxygen species–driven lipid peroxidation) (6, 9, 10). Both models imply that the unusual “leakiness” of endo- cytic compartments in DCs might not rely on any cell type–specific effectors, but on the regulation of ubiquitously expressed proteins through signaling (12, 13) and trafficking (14) events specific to cross-presenting cells. 1MRC Laboratory of Molecular Biology, Cambridge, UK. 2Department of Proteomics and Signal Transduction, Max Planck Institute of Biochemistry, Martinsried, Germany. 3UK Dementia Research Institute at the University of Cambridge, Department of Clinical Neurosciences, Cambridge, UK. *Corresponding author. Email: pkozik@mrc-lmb.cam.ac.uk †Present address: School of Biological Sciences, Faculty of Biology, Medicine and Health, Manchester Academic Health Science Centre, University of Manchester, Manchester, UK. Here, to identify DC-specific regulators of endocytic escape, we developed a flow cytometry–based assay to monitor escape in individual cells and applied it in a CRISPR- Cas9–based screen using cells specialized in cross-presentation, conventional DC1s (cDC1s). A CRISPR-Cas9 screen identifies Mpeg1 (perforin-2) as a regulator of endocytic escape To monitor endocytic escape, we used the 28-kDa type I ribosome-inactivating protein (RIP) saporin (5). Once in the cytosol, RIPs arrest translation through depurination of the sarcin-ricin loop in the 28S subunit of the ribosome (15). In contrast to type II RIPs, such as ricin, which comprise a domain that facil- itates entry into the cytosol, cytosolic delivery of type I RIPs is dependent on the cell-intrinsic efficiency of endosome-to-cytosol transport (5). To detect saporin-induced translation inhibi- tion, we monitored incorporation of puromycin, a structural analog of aminoacyl tRNAs, into nascent polypeptides (16). We labeled intra- cellular puromycylated polypeptides with a fluorescent 12D10 antibody allowing for flow cytometry–based readout of translation effi- ciency and thereby of endocytic escape (Fig. 1A). cDC1s, a subset of DCs that excels at cross- presentation in vivo (17), have been reported to have the most efficient endocytic escape pathway (17, 18). We thus developed the saporin-puromycin assay using a murine cDC1- like cell line, MutuDCs (13, 19, 20) (Fig. 1B and fig. S1, A and B). We confirmed that saporin escape is the rate-limiting step in the assay (fig. S1, C and D) and that the assay recapitulates previously reported differences in escape effi- ciency, namely, more efficient escape in cDC1s compared with cDC2s (fig. S2, A and C) (18, 21) and enhanced “leakiness” of endocytic com- partments in cells from lupus-prone MRL/ MpJ-Faslpr/J mice (fig. S2, B and D) (22). We then used the saporin-puromycin assay in a CRISPR-Cas9–based screen of 281 genes highly expressed in cDC1s compared with cDC2s (fig. S3). We used a mix of Atto550-labeled and unlabeled saporin, gated on cells with similar uptake efficiency, and sorted the cells into two bins: purolow (saporin escape, translation arrest) and purohigh (saporin retention, efficient trans- lation) (fig. S3C). In the absence of saporin, none of the single guide RNAs (sgRNAs) affected the translation rate (fig. S3E and table S1). The strongest hit in saporin-pulsed cells was Mpeg1 (Fig. 1C), with the four sgRNAs enriched in purolow versus purohigh populations (fig. S3F). Perforin-2 is necessary for endocytic escape in DCs Mpeg1 encodes perforin-2, a member of the membrane attack complex (MAC) and perforin superfamily (MACPF) of pore-forming proteins (23). Perforin-2 can form oligomeric pores on liposomes with an opening of at least 75 Å in diameter (24–26) (Fig. 1D). It was initially pro- posed that perforin-2 pores facilitate killing of intravacuolar bacteria (27), but these results were not replicated in a recent study (28). Here, we explored whether perforin-2 can function as an effector of endocytic escape. We generated Mpeg1KO MutuDCs and con- firmed protein depletion in sorted, sgRNA- expressing blue fluorescent protein (BFP) cells (fig. S4A). To account for the effect of passage numbers on MutuDC behavior (20), we cultured control cells expressing nontargeting (NT) sgRNA in parallel with the knockout (KO) line. We first demonstrated that disruption of Mpeg1 protected the cells from the cytotoxic effects of saporin and from death induced by a different RIP, gelonin (fig. S5A). We also tested the sensitivity of Mpeg1KO DCs to a glycopeptide chemotherapeutic, bleomycin A2, which induces DNA damage, but owing to its hydrophilicity does not enter the cells efficiently (29). We used automated imaging to monitor MutuDC growth rate and found that loss of perforin-2 rendered the cells more resistant to bleomycin- mediated cytotoxicity (fig. S5A). Notably, the Mpeg1KO cells were not protected against the effects of poly(I:C), which induces cell death via endosomal Toll-like receptor 3 (TLR3) (30), or against membrane-permeable cycloheximide (fig. S5A). We next verified that the sensitivity of perforin-2–expressing DCs to cytosolic toxins is due to efficient escape rather than due to efficient uptake. In the saporin-puromycin assay with Atto550-labeled saporin, endocytic escape was impaired in Mpeg1KO DCs, even though they internalized similar amounts of saporin compared to NT cells (Fig. 1E). The differences in escape were also not due to changes in DC activation, because neither NT nor Mpeg1KO MutuDCs were activated by saporin (fig. S5B). Finally, we rescued saporin Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Saporin-puromycin assay to monitor endocytic escape in DCs. (A and B) Saporin-puromycin assay. (A) Schematic representation. Saporin- pulsed cells are labeled with puromycin to monitor translation rate. Puromycin incorporated into nascent peptides is detected with an aPuromycin Ab and flow cytometry. If saporin is retained within the endosomes, translation remains high. When saporin escapes into the cytosol, it depurinates ribosomes, inducing translation arrest. (B) Representative flow cytometry plot. MutuDCs were incubated with 0.5 mg/ml of saporin followed by 0.01 mg/ml puromycin (purple histogram), with puromycin alone (yellow), or in media only (gray). Cells in translation arrest are denoted by the purple, gate. See also fig. S1B. (C) Volcano plots showing the sgRNA enrichment analysis for the saporin- puromycin endocytic escape screen. Each of the dots represents one targeted gene. Data represent the combined mean enrichment scores and the non- adjusted P values from three independent experiments (Fisher’s method). See also fig. S3. (D) Schematic representation of the different perforin-2 domains alongside structures of single-subunit (PDB ID: 6U2K) and the hexadecameric perforin-2 pore (PDB ID: 6SB5). (E) Quantification of translation arrest in Mpeg1KO and NT MutuDCs. Cells were pulsed with saporin (11:1 unlabeled: Atto550-labeled saporin) for 2 hours, and translation was monitored by a 30-min puromycin chase. The x axis represents Atto550 mean fluorescence intensity (MFI) normalized to the NT MutuDC Atto550 MFI at the highest saporin concentration. Data represent mean and SEM of three independent experiments; ns, not significant; **P < 0.01 using a multiple unpaired t test (two-stage step-up, Benjamini, Krieger, and Yekutieli). Significance symbols in the plot refer to the differences in proportion of cells in translation arrest. Differences in saporin Atto550 MFI were not significant. See also fig. S4A. (F) Mpeg1KO MutuDCs were reconstituted with the indicated Mpeg1 mutants or with mScarlet only and used in the saporin-puromycin assay with a 2-hour pulse with 0.1 mg/ml saporin. Data are representative of three independent experiments. See also fig. S4C. import without affecting uptake by transducing Mpeg1KO cells with full-length sgRNA-resistant Mpeg1 (figs. S4C and S5C). Next, we addressed whether the pore-forming ability of perforin-2 is required for endocytic escape. Perforin-2 forms pores by oligomeriza- tion and unwinding of two helices in the MACPF domain into b sheets (shown in red in Fig. 1D) (24, 25). To test whether this conforma- tional change is required for endocytic escape, we generated two mutants: Mpeg1G212V/A213V, with mutations in the conserved MACPF motif (31), and Mpeg1K251C/G286C, with a disulfide bond known to constrain one of the pore-forming helices preventing pore formation in vitro (24). Both Mpeg1G212V/A213V and Mpeg1K251C/G286C were expressed at levels similar to those of Mpeg1WT, but neither rescued saporin escape in Mpeg1KO MutuDCs (Fig. 1F and fig. S4C). Thus, perforin-2 pores mediate endocytic escape in cross-presenting DCs. Perforin-2 is sufficient for endocytic escape in nonimmune cells Because perforin-2 expression is restricted to antigen-presenting cells (32, 33), we asked whether it is sufficient for endocytic escape in nonimmune cells. We generated human em- bryonic kidney 293T (HEK293T) and HeLa cells coexpressing murine Mpeg1 with mScarlet or BFP, or expressing the fluorescent protein alone, and used them in a range of escape assays. Ectopic expression of perforin-2 was suffi- cient to promote saporin escape in HEK293Ts (Fig. 2A). We also monitored escape of 34-kDa b-lactamase using a cytosolic dye CCF4 (34, 19). CCF4 consists of fluorescein and 7-hydroxycou- marin linked by a b-lactam ring, which is cleaved when b-lactamase escapes into the cytosol, resulting in a shift in fluorescence emission (Fig. 2B). Expression of perforin-2 in HeLa cells increased the frequency of cells with cleaved CCF4 and thus the efficiency of b-lactamase escape (Fig. 2B and fig. S6). We then adopted a split luciferase-based assay to monitor endocytic escape of microtubule- associated protein, tau (35). We expressed the large 18-kDa NanoLuc subunit (LgBiT) in the cytosol of HEK293Ts and pulsed the cells with oligomers of 42-kDa tau fused to the short HiBiT peptide. Upon entry into the cytosol, LgBiT binds HiBiT, resulting in catalytically active luciferase and luminescence in the pres- ence of the Nano-Glo(R) substrate. Again, in perforin-2–expressing HEK293Ts, tau-HiBit escaped more efficiently compared with the Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Perforin-2 is sufficient for endocytic escape of cargo in non-immune cells. (A) HEK293Ts and Mpeg1-complemented HEK293Ts were pulsed with saporin (11:1 unlabeled:Atto550-labeled saporin) for 2 hours, and translation was monitored by a 30-min puromycin chase. The x axis represents Atto550 MFI normalized to the WT cells pulsed with 0.5 mg/ml saporin. Data represent mean and SEM of three independent experiments; ns, not significant; ****P < 0.0001 using a multiple t test (Bonferroni-Dunn). Significance symbols on the plot refer to the differences in cells in translation arrest. Differences in saporin Atto550 MFI were not significant. (B) Cytosolic escape of b-lactamase in cells loaded with the CCF4 results in CCF4 cleavage, loss of fluorescence resonance energy transfer, and shift in emission fluorescence. HeLa cells expressing either Mpeg1IRES-mScarlet or mScarlet only were pulsed with b-lactamase for the indicated time. b-lactamase escape was monitored by measuring the shift in fluorescence emission by flow cytometry. Data represent mean and SEM of three independent experiments, ns, not significant; ns, not significant; **P < 0.01; ***P < 0.001 using a multiple t test (two-stage step-up, Benjamini, Krieger, and Yekutieli) For gating strategy, see fig. S6. (C) To monitor the escape of tau oligomers, cells expressing NLS-eGFP- LargeBiT were pulsed with Tau-HiBit oligomers. The escape of Tau-HiBiT into the cytoplasm allows binding to the 18-kDA luciferase subunit, LgBiT. This results in reconstitution of catalytic activity and generation of luminescence. NLS-eGFP- LargeBiT HEK293Ts expressing either Mpeg1IRES-BFP or BFP only were pulsed with tau-HiBiT for the indicated time. Following substrate addition, luminescence and cell viability were assessed. Relative luminescent units (RLUs) were then normalized to viability per well. Data represent mean and SEM of three independent experiments each with six technical replicates, ***P < 0.001 using a paired t test. (D) HEK293Ts were plated in the presence or absence of bleomycin and cultured in an Incucyte for 48 hours to monitor the growth rate. Data represent mean and SEM of three independent experiments each with four wells per condition; ns, not significant; *P < 0.5; **P < 0.01; ****P < 0.0001 using a multiple t test (Bonferroni-Dunn). control line (Fig. 2C). Finally, we demonstrated that expression of perforin-2 renders HeLa cells more sensitive to bleomycin-mediated toxicity (Fig. 2D). Thus, ectopic expression of perforin-2 is sufficient to drive endocytic es- cape in nonimmune cells. Perforin-2 is proteolytically processed in lysosomes Given the cytotoxic potential of pore-forming proteins, we asked how cDC1s regulate pore formation and restrict it to antigen-containing compartments (19, 35). Perforin-2 is the only known mammalian pore-forming protein with an additional transmembrane domain (TMD) that is not involved in pore formation (Fig. 1D). The TMD has been proposed to act as an anchor preventing damage to endogenous membranes and orientating the pore toward intravacuolar bacteria (24, 27). We hypothe- sized that the ectodomain would have to be proteolytically released to facilitate endocytic escape. To investigate the proteolytic processing of perforin-2, we analyzed a SILAC (stable isotope labeling by amino acids in cell culture)–based organellar mapping dataset generated previ- ously (19). The maps were prepared by mass spectrometry–based analysis of the fractionated postnuclear supernatants from MutuDCs (fig. S7A) (36). In the original maps, perforin-2 had a profile similar to those of lysosomal proteins. Here, instead of analyzing protein profiles, we analyzed each tryptic peptide individually, only including endo- or lysosomal proteins (Fig. 3, A and B). Peptides derived from endosome- and lysosome-resident proteins formed separate clusters, as did peptides derived from trans- membrane and luminal proteins in the lyso- some cluster. Fifteen out of 17 perforin-2– derived peptides clustered with soluble rather than transmembrane lysosomal proteins, con- sistent with the hypothesis that the perforin-2 ectodomain is released from the TMD anchor. The remaining perforin-2 peptides, p340-372 [within the epidermal growth factor (EGF)– like domain] and p629-635 (within the TMD- proximal region), coclustered with endosomal proteins (such as Vps35, Fig. 3A), suggesting that they are present in the full-length protein as it transits through endosomes but are absent (cleaved) once perforin-2 reaches lysosomes. We confirmed that perforin-2 was proteo- lytically processed using antibodies against the MACPF, P2, and the C-terminal tail (C-term) (fig. S7, B and C). All three antibodies detected full-length perforin-2 at 70 kDa. We also iden- tified a 40-kDa MACPF and a 30-kDa P2 frag- ments, consistent with the cleavage in the EGF domain. The aC-term antibody detected only the full-length protein, indicating that the tail (and most likely the TMD) are rapidly degraded after release of the ectodomain. Bafilomycin Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Perforin-2 undergoes proteolytic cleavage, releasing its pore- forming domain into the organellar lumen. (A and B) (A) Principal-component analysis of mass spectrometry–based organellar mapping of MutuDCs (19) (see also fig. S7A). The maps were prepared from control MutuDCs and cells treated with drugs that promote lysosomal leakiness, prazosin and tamoxifen. Peptides derived from lysosomal and endosomal proteins are represented by filled and empty circles, respectively. The different colors indicate localization of the protein within the corresponding organelle. Perforin-2 peptides are displayed as filled black circles regardless of their localization. (B) Mapping of the different lysosomal and endosomal perforin-2 peptides detected by organellar mass spectrometry onto the different perforin-2 domains and structure. (C) Perforin-2 levels in NT MutuDCs treated with 0.5 mM BafA1 for 3 hours were assessed by immunoblot under reducing conditions using the aMACPF and aC-terminal tail antibodies. (D) Confocal microscopy images of MutuDCs stained for perforin-2 with either aC-terminal tail or aMACPF antibodies (red), Vps35 (green), and lysotracker (blue). Data represent two independent experiments each with at least 80 cells, ***P < 0.0001 using a Kolmogorov- Smirnov test. (E) BafA1- and CpG-induced changes in the abundance of tryptic and semi-tryptic (cleaved) perforin-2 peptides. Control and treated cells (1 mM BafA1 or 1 mM CpG) were analyzed by mass spectrometry, and peptide intensities were normalized to the corresponding protein intensities. Statistical analysis was performed with a two-sided student’s t test. P values (y axis) and fold change in abundance (line thickness) in treated versus control cells are shown. The amino acid position indicates the location of the peptides along the different perforin-2 domains. (F) Differences in the abundance of tryptic and semi-tryptic (cleaved) perforin-2 peptides between control and AEPKO MutuDCs (fig. S9B) by full proteome mass spectrometry. The analysis was performed as in (E). Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Perforin-2 undergoes pH-dependent maturation in antigen-containing phagosomes. (A) Schematic representation of the phagoFACS assay. Cells are pulsed with OVA-beads and allowed to internalize them. After an indicated chase period, noninternalized beads are labeled with an aOvalbumin antibody. After cell homogenization, phagosomes are stained with antibodies against OVA coupled to an alternative fluorophore and phagosomal markers. (B) Mpeg1KO and NT MutuDCs were pulsed with OVA- beads and chased for the indicated time in the presence of either BFA or BafA1. Isolated phagosomes were stained with antibodies against Lamp1 and either the perforin-2 C-terminal tail (top panel) or MACPF domain (bottom panel). Data are representative of three independent experiments. See also fig. S11 for gating strategy, quantification, and additional plots. A1, a vacuolar-type ATPase (V-ATPase) inhib- itor that interferes with lysosomal acidifi- cation, accumulated full-length perforin-2, consistent with the hypothesis that proteo- lytic processing occurs in lysosomes (Fig. 3C). Finally, confocal microscopy confirmed that the aC-term antibody, which we predicted to recognize the immature (full-length) perforin-2, colocalized with Vps35 (Fig. 3D), and the aMACPF antibody colocalized with lyso- tracker, but not with Vps35. This result, together with the organellar mapping data, suggests that most of the protein is in lysosomes at steady-state. Thus, full-length perforin-2 resides in (or transits through) endosomes, and upon reaching low pH compartments, undergoes maturation involving at least two cleavage events. Perforin-2 maturation is controlled by asparagine endopeptidase (AEP) We then asked whether perforin-2 maturation might be regulated by antigen-associated sig- nals. From a panel of TLR agonists, CpG (TLR9- agonist) and Toxoplasma profilin (TLR11-agonist) were most efficient in promoting perforin-2 proteolytic processing (fig. S8). To further characterize the putative cleavage sites, we performed comparative proteomics analysis of tryptic and semi-tryptic peptides from CpG-, BafA1-, and mock-treated cells (Fig. 3E). Semi- tryptic peptides are considered to be derived from proteins that were cleaved in the cell, before sample processing. The semi-tryptic peptide p340-349 (within the EGF domain) was enriched in CpG-treated cells and depleted in BafA1-treated cells, whereas tryptic peptides near the TMD, p621-628 and p629-637, were depleted in CpG-treated cells and enriched in BafA1-treated cells. Several of the nontryptic peptides in the EGF region terminated on asparagine (fig. S9A), suggesting that perforin-2 processing might be mediated by AEP (37). Indeed, the cleavage pattern in the EGF domain was dif- ferent in AEPKO MutuDCs (fig. S9B) compared with control cells (see peptides p340-358 and p340-351, Fig. 3F), suggesting that AEP medi- ates perforin-2 maturation, but its activity can be replaced by other enzymes [most likely cathepsins (38)]. Immunoblot analysis con- firmed that the EGF cleavage is less efficient (albeit not completely abolished) in the absence of AEP, resulting in accumulation of the 60-kDa ectodomain (fig. S9C). We were also able to reconstitute this cleavage in vitro using re- combinant perforin-2 and AEP (fig. S9D). In summary, perforin-2 maturation is con- trolled by at least two cleavage events, both oc- curring at steady-state but stimulated by TLR signaling. The cleavage in the TMD-proximal CTT domain releases the ectodomain into the lysosomal lumen to orient the pore-forming domain toward the endogenous membranes. Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. Antigen cross-priming is impaired in vivo in the absence of perforin-2. (A) Perforin-2 expression was assessed by intra- cellular staining with an aPerforin-2 antibody and flow cytometry in Mpeg1+/+ and Mpeg1−/− spleno- cytes. cDC1s are defined as Lineage (CD3, CD19, NK1.1)−, F4/80−, CD11c, XCR1+; cDC2s as Lineage (CD3, CD19, NK1.1)−, F4/80−, CD11c, CD172a+; and pDCs as Lineage (CD3, CD19, NK1.1)−, F4/80-, CD11cintSiglecH+. For gat- ing strategies, see fig. S13C. (B) CD11c+ magnetically enriched splenocytes from WT and Mpeg1−/− mice were pulsed with saporin for 2 hours, and translation was monitored by a 30-min puromycin chase. cDC1s are defined as CD11c+, XCR1+, cDC2s are defined as CD11c+, CD172a+, and pDCs are defined as CD11cintSiglecH+. Dot plots are representative of two independent experiments. For gat- ing strategies, see fig. S14E. (C and D) Wild-type and Mpeg1−/− mice were intravenously (i.v.) injected with 0.5 × 106 cell trace violet (CTV)–labeled magnetically purified OT-I cells. One day later, mice were injected i.v. with 1 × 106 UVC- irradiated (240 mJ/cm2) 3T3 cells, coated with 10 mg/ml ovalbumin as antigen source and 0.25 mg/ml Poly(I:C) as an adjuvant. Three days later, OT-I proliferation was assessed by flow cytometry (C). OT-I are defined as Lineage (CD19, F4/80, CD11c)−, CD3+, CD4−CD8+, TCRvb5.1, 5.2+TCRva2+, CTV+. For gating strategy, see fig. S15A. (D) Normalized OT-I counts 3 days after intravenous antigen injection. Each dot corresponds to an indi- vidual mouse, with three to five mice per group. For each experi- ment, OT-I counts per 1 × 106 splenocytes were normalized to the average of WT controls. Data represent five independent experiments. ns, not significant; **P < 0.01 using an unpaired t test. The AEP-mediated cleavage in the EGF region, which is not required for pore formation in vitro (24, 39), may provide additional flexibility to complete pore insertion in vivo or may serve to inactivate the pores. Perforin-2 undergoes maturation in antigen-containing compartments To test whether perforin-2 undergoes matura- tion upon recruitment to antigen-containing compartments or whether antigen-containing compartments acquire mature perforin-2, we analyzed perforin-2 processing in phagosomes. We first confirmed that perforin-2 mediates the escape of bead-conjugated saporin from phagosomes into the cytosol (fig. S10). We then followed perforin-2 maturation in individual phagosomes by phagoFACS, a technique in which DCs are pulsed with ovalbumin (OVA)–coated beads and the resulting phagosomes are an- alyzed by flow cytometry (Fig. 4A). Using the aC-term antibody, we demonstrated that perforin-2 was rapidly recruited to phagosomes, reaching its highest levels within 30 min (Fig. 4B and fig. S11, A and B). This rapid acquisition suggests that perforin-2 is recruited before phagosome-lysosome fusion (40). Indeed, bre- feldin A (BFA), an inhibitor of ARF GTPases that blocks protein trafficking through the early secretory pathway, inhibited perforin-2 (but not Lamp1) recruitment, consistent with the delivery of full-length perforin-2 to phago- somes from the Golgi (Fig. 4B and fig. S11, B Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E and C). In phagoFACS experiments with the aMACPF antibody, perforin-2 acquisition was also BFA dependent. However, the aMACPF staining was restricted to Lamp1+ phagosomes and increased over time, whereas the aC-term signal was gradually lost (Fig. 4B and fig. S11, A and B). Thus, during phagosome maturation, the C-terminal part of perforin-2 is cleaved off and degraded, whereas the ectodomain under- goes a conformational change that exposes an epitope recognized by the aMACPF antibody. To confirm that perforin-2 maturation in phagosomes is pH dependent, we first asked whether MutuDC phagosomes acidify given the conflicting reports on whether the pH in DC phagosomes remains high (41–43) or de- creases (44–46) over time. Using pHrodo-beads, we found that phagosome acidification in MutuDCs was abrupt and occurred on a time- scale comparable to that of Lamp1 acquisition (fig. S12). Inhibition of phagosomal acidifica- tion with BafA1 interfered with both gradual loss of the aC-term signal and acquisition of the aMACPF signal (Fig. 4B and fig. S11B), confirming that perforin-2 maturation is regulated by pH. The pHrodo experiments also revealed that phagosomal pH was similar in NT and Mpeg1KO MutuDC (fig. S12), suggesting that perforin-2 pores are either not permissive to protons, pore formation is transient, or the cells can compen- sate for proton loss. Furthermore, these data suggest that membrane integrity of phagosomes is not compromised during perforin-2–mediated escape, in line with the observation that phago- somes containing OVA-beads do not recruit galectin-3, a marker of damaged compartments (6). Consistent with these data, OVA degra- dation was not affected by knocking out Mpeg1, confirming that perforin-2 does not drastically alter the degradative potential of phagosomes (fig. S11, D and E). Thus, perforin-2 undergoes pH-dependent maturation in antigen-containing compartments, and perforin-2–mediated endocytic escape of antigens can occur while preserving the overall integrity of the phagosomal membrane. Perforin-2 is expressed in antigen-presenting cells and facilitates cross-presentation To test whether perforin-2 is involved in the delivery of antigens for cross-presentation, we generated Mpeg1−/− mice (fig. S13A and B). Notably, knocking out Mpeg1 did not result in any obvious disease phenotype or a change in immune cell frequencies (fig. S13, C and D). We used intracellular staining to confirm that perforin-2 expression is restricted to splenic cDC1s (Lin−CD11c+XCR1+), as well as other cell types previously shown to cross-present, includ- ing splenic macrophages (Lin−F4/80+), plasma- cytoid DC (pDCs, Lin−F4/80−CD11cintSiglecH+), and Ly6C+ monocytes (Fig. 5A and fig. S14A) (47–50). In the spleen, only a small, CX3CR1+ subpopulation of cDC2s expressed perforin-2, whereas in the lungs, a large fraction of cDC2s was perforin-2 positive (fig. S14, B and C). Consistent with perforin-2 expression patterns, knocking out Mpeg1 decreased the efficiency of endocytic escape in splenic cDC1s and pDCs, but not in cDC2s (Fig. 5B). We also asked whether perforin-2–mediated escape delivers antigens for cross-presentation. To assess the efficiency of antigen presentation in vivo, we adoptively transferred OT-I T cells (expressing a T cell receptor specific to H2- Kb MHC-I with OVA257-264 peptide) and mo- nitored OT-I proliferation after immunization. Because cross-presentation of soluble OVA, which can be processed by cell types other than cDC1s (48, 50), was not significantly affected in the Mpeg1−/− mice (fig. S15), we used dead cells as an antigen source. Compared to wild-type (WT) mice, Mpeg1−/− had fewer OT-I T cells after immunization with ultraviolet C (UVC)– irradiated, OVA-coated fibroblasts (Fig. 5C). Similarly, Mpeg1−/− bone marrow-derived cDC1s [from FMS-like tyrosine kinase 3 ligand (Flt3 L) and granulocyte-macrophage colony-stimulating factor (GM-CSF) cultures] displayed impaired endocytic escape in the saporin assay and a reduced capacity to cross-present cell-associated antigens (fig. S16). Thus, loss of perforin-2 leads to a defect in cross-presentation of cell-associated antigens in vitro and in vivo, suggesting that in cross-presenting cells, endocytic pores provide a route for cytosolic entry of antigens. Conclusions Here, we uncovered a mechanism of endocytic escape that is governed by a cell type–specific pore-forming effector protein, perforin-2. The role of perforin-2 in the delivery of antigens for cross-presentation suggests that the immune system evolved to use two related pore-forming effectors, perforin-1 and perforin-2, during diff- erent stages of adaptive immune responses. Perforin-1, expressed by cytotoxic T cells, is a well-characterized effector used for the deliv- ery of granzymes into the cytosol of target cells (51). Our data suggest that perforin-2 delivers endocytic contents into the cytosol of cross- presenting DCs to enable generation of MHC- I:peptide complexes and T cell priming. We do not exclude the possibility that in different contexts, membrane destabilization may also result in antigen delivery for cross- presentation (6, 9, 52). Membrane integrity is regulated by a wide range of mechanisms (53, 54), and different perturbations can result in leakage of endosomal contents into the cytosol under pathological conditions. Perforin 2–mediated endocytic escape, however, appears to occur without compromising the overall stability of the endocytic compartments. The restricted expression of Mpeg1 to sub- sets of professional antigen-presenting cells points to an important role of perforin-2 during the initiation of immune responses. The ability to genetically manipulate perforin-2–mediated endocytic escape provides a tool for exploring the contribution of the escape pathway to anticancer and antiviral immunity in vivo. REFERENCES AND NOTES 1. M. J. Palmowski et al., J. Immunol. 177, 983–990 (2006). J. Helft et al., J. Clin. Invest. 122, 4037–4047 (2012). 2. 3. D. J. Theisen et al., Science 362, 694–699 (2018). 4. E. Giampazolias et al., Cell 184, 4016–4031.e22 (2021). 5. M. Kovacsovics-Bankowski, K. L. Rock, Science 267, 243–246 (1995). 6. 7. M. Zehner et al., Immunity 42, 850–863 (2015). 8. M. Zehner et al., Proc. Natl. Acad. Sci. U.S.A. 108, 9933–9938 J. Canton et al., Nat. Immunol. 22, 140–153 (2021). (2011). I. Dingjan et al., Sci. Rep. 6, 22064 (2016). 9. 10. S. C. Nalle et al., PLOS ONE 15, e0238484 (2020). 11. J. E. Grotzke et al., Proc. Natl. Acad. Sci. U.S.A. 114, E5910–E5919 (2017). 12. Y. Ding et al., Nat. Immunol. 17, 1167–1175 (2016). 13. J. Canton et al., Nat. Immunol. 22, 140–153 (2021). 14. S. Zelenay et al., J. Clin. Invest. 122, 1615–1627 (2012). 15. K. Nielsen, R. S. Boston, Annu. Rev. Plant Physiol. Plant Mol. Biol. 52, 785–816 (2001). 16. E. K. Schmidt, G. Clavarino, M. Ceppi, P. Pierre, Nat. Methods 6, 275–277 (2009). 17. K. Hildner et al., Science 322, 1097–1100 (2008). 18. M. L. Lin et al., Proc. Natl. Acad. Sci. U.S.A. 105, 3029–3034 (2008). 19. P. Kozik et al., Cell Rep. 32, 107905 (2020). 20. S. A. Fuertes Marraco et al., Front. Immunol. 3, 331 (2012). 21. M. Gros et al., Cell Rep. 40, 111205 (2022). 22. A. J. Monteith et al., Proc. Natl. Acad. Sci. U.S.A. 113, E2142–E2151 (2016). 23. R. McCormack, L. R. de Armas, M. Shiratsuchi, J. E. Ramos, E. R. Podack, J. Innate Immun. 5, 185–194 (2013). 24. T. Ni et al., Sci. Adv. 6, eaax8286 (2020). 25. S. S. Pang et al., Nat. Commun. 10, 4288 (2019). 26. X. Yu, T. Ni, G. Munson, P. Zhang, R. J. C. Gilbert, EMBO J. 41, e111857 (2022). 27. R. M. McCormack et al., eLife 4, e06508 (2015). 28. S. Ebrahimnezhaddarzi et al., Immunol. Cell Biol. 100, 529–546 (2022). 29. M. J. Jaroszeski et al., Anticancer Drugs 11, 201–208 (2000). 30. S. A. Fuertes Marraco et al., PLOS ONE 6, e20189 (2011). 31. C. J. Rosado et al., Science 317, 1548–1551 (2007). 32. T. S. Heng et al., Nat. Immunol. 9, 1091–1094 (2008). 33. J. Lonsdale et al., Nat. Genet. 45, 580–585 (2013). 34. I. Cebrian et al., Cell 147, 1355–1368 (2011). 35. B. J. Tuck et al., Cell Rep. 39, 110776 (2022). 36. D. N. Itzhak, S. Tyanova, J. Cox, G. H. Borner, eLife 5, e16950 (2016). 37. J.-M. Chen, P. M. Dando, R. A. E. Stevens, M. Fortunato, A. J. Barrett, Biochem. J. 335, 111–117 (1998). 38. S. E. Ewald et al., J. Exp. Med. 208, 643–651 (2011). 39. F. Jiao et al., Nat. Commun. 13, 5039 (2022). 40. F. M. Cruz, J. D. Colbert, K. L. Rock, EMBO J. 39, e102020 (2020). 41. A. Savina et al., Immunity 30, 544–555 (2009). 42. A. R. Mantegazza et al., Blood 112, 4712–4722 (2008). 43. A. Savina et al., Cell 126, 205–218 (2006). 44. S. C. Khouili et al., Cell Rep. 33, 108468 (2020). 45. J. M. Rybicka, D. R. Balce, S. Chaudhuri, E. R. O. Allan, R. M. Yates, EMBO J. 31, 932–944 (2012). 46. M. Samie, P. Cresswell, Nat. Immunol. 16, 729–736 (2015). 47. C. Bosteels et al., Immunity 52, 1039–1056.e9 (2020). 48. M. Enders et al., J. Immunol. 204, 87–100 (2020). 49. M. Oberkampf et al., Nat. Commun. 9, 2241 (2018). 50. N. M. Kretzer et al., J. Exp. Med. 213, 2871–2883 (2016). I. Voskoboinik, J. C. Whisstock, J. A. Trapani, Nat. Rev. 51. Immunol. 15, 388–400 (2015). 52. P. Cresswell, Immunogenetics 71, 141–160 (2019). 53. M. L. Skowyra, P. H. Schlesinger, T. V. Naismith, P. I. Hanson, Science 360, eaar5078 (2018). 54. I. Maejima et al., EMBO J. 32, 2336–2347 (2013). AC KNOWLED GME NTS We thank members of the Kozik laboratory for discussions and feedback regarding this work; we also thank G. M. Griffiths and L. C. James for reading of the manuscript and their feedback. We are indebted to N. Hacohen and T. Eisenhauer for their help setting up CRISPR-Cas9 screens and G. Slodkowicz for help with statistical Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E analysis. We also thank A. Crisp for help with data processing. We are grateful to M. Mann for his support and I. Paron and T. Heymann at the Max Planck Institute of Biochemistry for technical assistance with mass spectrometry. Finally, we thank the animal facility/ARES staff, genotyping facility, flow cytometry core, and microscopy core at the MRC Laboratory of Molecular Biology for their technical assistance. Funding: This work was supported by the UK Medical Research Council (MRC grant no. MC_UP_1201/ 26). P.R.-S. was supported by an MRC CellTech Research Fellowship (MRF-104-0007-S-RODRI). P.A.K. was supported by the Boehringer Ingelheim Fonds PhD Fellowship. G.H.H.B., A.K.D., and J.P.S. were funded by the Max Planck Society for the Advancement of Science. B.J.T. was supported by the Cambridge Trust Vice Chancellor Award and Hughes Hall Edwin Leong PhD scholarship. W.A.M. is a Lister Institute Fellow and supported by a Sir Henry Dale Fellowship jointly funded by the Wellcome Trust and the Royal Society (206248/Z/17/Z). W.A.M. was further supported by the UK Dementia Research Institute, which receives its funding from DRI Ltd, funded by the UK Medical Research Council, Alzheimer’s Society, and Alzheimer’s Research UK. Author contributions: P.R.S. designed and performed experiments and wrote the manuscript. M.L. designed and performed experiments and contributed to the manuscript. P.A.K. performed experiments and provided scientific insight. A.K.D., J.P.S., and G.H.B. carried out mass spectrometry experiments and proteomics data analysis. R.P. performed experiments. B.J.T. and W.A.M. carried out Tau-HiBit assay. P.K. supervised the study, designed and performed experiments, and wrote the manuscript. Correspondence and requests for materials should be addressed to Patrycja Kozik (pkozik@mrc-lmb.cam.ac.uk). Competing interests: The authors declare no competing interests. Data and materials availability: All data are available in the main text or the supplementary materials. Data are available via ProteomeXchange with identifier PXD041861. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adg8802 Materials and Methods Figs. S1 to S16 Tables S1 to S5 References (55–61) MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. Submitted 27 January 2023; accepted 3 May 2023 10.1126/science.adg8802 Rodríguez-Silvestre et al., Science 380, 1258–1265 (2023) 23 June 2023 8 of 8
10.1126_science.adh1331
RES EARCH GEOPHYSICS Tremor signals during fluid injection are generated by fault slip Shankho Niyogi1*, Abhijit Ghosh1, Abhash Kumar2, Richard W. Hammack2 Seismic tremor signals, also known as long-period, long-duration signals, have been reported in several locations where fluid injection for enhanced oil and gas exploration is taking place. However, the origin of these signals remains poorly constrained. We studied seismic tremor signals in Wellington Field, Kansas, using a seismic array during a carbon dioxide injection program. We show that these signals are generated below the surface during the time of carbon dioxide injection. They have a distinct spectral signature, similar to those observed in glacial and volcanic environments. The tremor sources are located near the injection site and aligned with preexisting faults. Modeling results imply that such tremors are generated by frictional slip on fault. These observations may reveal an important deformation mode, which is useful for studying associated stress, seismicity, and triggering, as well as for tracking fault activities during injection operations of all fluids, including supercritical carbon dioxide. C O2 injection in the subsurface can be performed both for sequestration and for enhanced oil recovery purposes. It is a discipline that has recently witnessed growth because of increasing efforts to mitigate global warming and growing energy demand. When CO2 is injected into an oil reser- voir, it reduces the viscosity of the residual oil and expands its volume, facilitating extraction, extending the productive life of the hydrocarbon field, and removing the injected CO2 out of the atmosphere. However, the injection of these fluids can increase the pore pressure in the subsurface and potentially cause critically stressed faults near the injection site to slip. When seismic events such as earthquakes are generated from such a mechanism, this is known as induced seismicity. In contrast to natural earthquakes, some induced seismic events can show very different signal characteristics. Such signals are typically reported as long-period, long-duration events in several fluid injection sites (1–6). Physical mechanisms of such events, their underlying physics, and their relation to the injected fluid, however, remain enigmatic. The waveform characteristics of these long- period, long-duration signals are similar to the tectonic tremors found in large natural faults. Contemporary studies report that the pres- ence of fluids in faults and subduction zones is one of the factors behind the origin of tectonic tremors and is largely driven by aseismic slip (7–9). Laboratory studies have shown that slow injection rates can accelerate aseismic creep while maintaining stable slip due to the pressur- ized zone being confined within a critical size for an unstable slip in fault gouge samples (10). Similarly, aseismic deformation was found to be the dominant response when fluid injections were performed in natural fault systems in the 1University of California, Riverside, CA, USA. 2National Energy Technology Laboratory, Pittsburgh, PA, USA. *Corresponding author. Email: sniyo001@ucr.edu Low Noise Underground Laboratory [Labora- toire Souterrain Bas Bruit (LSBB)] of Rustrel, France, to induce slip (11). We also have evi- dence of fluid injection–induced earthquakes exhibiting broader body wave pulses and lower frequency coda (12), and these characteristics have been attributed to slower rupture speeds and lower stress drop values, which makes them similar to low-frequency earthquakes with regard to source characteristics. We observed low-frequency, long-duration, emergent signals in the seismic monitoring sta- tions [Incorporated Research Institutions for Seismology (IRIS) network identifier: ZA] placed around the injection well 2-32 in Wellington Field, Kansas, during a supercritical CO2 injec- tion program (6) (Fig. 1). The supercritical CO2 was injected into the Mississippian carbonate reservoir in a fluid state (6). Additionally, waste- water was being disposed of in the deeper, aquitard-bounded Arbuckle Group Formation. On the basis of our analysis of the United States Geological Survey (USGS) earthquake catalogs, the onset of the injection program did not cause an increase in local earthquakes, especially around the injection well. Most earthquakes in the south- central Kansas region are caused by oil and gas activities such as wastewater disposal from both local areas and from northern Oklahoma caused by far-field pressure diffusion (13). The locations of these earthquakes show that they are aligned with the structural features of this region, like the Humboldt Fault Zone and the Central Kansas Uplift (Fig. 1). Historical earth- quakes in Kansas have also taken place on the faults along these features. The Mississippian Group of rocks and the Cambro-Ordovician Arbuckle Group in Wellington Field are the tar- get lithology for CO2-enhanced oil recovery and storage, respectively (14). Exploration seismic data show that both of these formations are con- nected to the basement faults in Wellington Field (14). Initial studies of these low-frequency events suggested aseismic expression of these fault movements (6). We focused on the coherency and spectral signatures of these events, which make them similar to harmonic tremors reported earlier from volcanic and glacial settings. We explain the signals using physical models derived from these sources. For simplicity, we refer to these long-period, long-duration events as tremors. Investigation of unknown seismic signals often involves a separate analysis for looking into possible anthropogenic sources. We con- ducted independent analysis of these events with regard to their origin from freight trains, vehicular traffic, aerial sources such as helicop- ters, and machinery equipment such as pumps and generators on the surface near the seismic array. The presence of equipment does create noise, which is shown in the clearly visible parts of the spectrograms. However, a number of events with high signal-to-noise ratio clearly establishes the characteristics of the tremor signal. We have used that to make a tremor catalog based upon these well-defined waveform characteristics. A notable characteristic of these tremor events are the lines gliding across frequencies between 1 and 5 Hz observed in the spectrograms. For many of the events displaying similar spectral characteristics with gliding spectral lines, slow- ness and back azimuth could not be robustly determined because of the inherent noise con- tent. Therefore, we did not include them in our final analyses. Observations of such kinds of low-frequency tremor events, their source lo- cations, and their spatiotemporal distribution may provide a useful and benign way of track- ing the migration of injected fluids in the subsurface. Using an array response function with a synthetic waveform showed that the array could perform reasonably well at 1 to 5 Hz, the dominant frequency band of interest, and up to at least 0.3 s/km, which was suitable for our analysis. Background The Arbuckle Formation consists mainly of interlayered dolomites and carbonates and was deposited in a shallow epicontinental sea in Kansas and Oklahoma during the late Cam- brian to the Middle Ordovician. Karst formations within it created permeable zones for waste- water disposal, sealed by the overlying Devonian Woodford shale formation. The Mississippian limestone formations were later deposited in similar shallow marine environment settings and are bounded by unconformities. Oil deposits of the Mississippian formations are mostly a combination of structural-stratigraphic traps where the present-day CO2 injection is going on. Tectonic events such as the Nemaha Uplift and the Midcontinent Rift event influenced the structure of the Mississippian and Arbuckle Formation, creating deep vertical faults in the subsurface. Subsurface formations in southern Kansas were strongly affected by these tectonic Niyogi et al., Science 381, 553–558 (2023) 4 August 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E events (15–23), and this is where most earth- quake activity has occurred, including an esti- mated magnitude 5.1 event in 1867. Information about subsurface structures in Wellington Field comes from three-dimensional seismic reflection data and formation micro- imaging (FMI) well logs (14). Analyzing the seismic data shows mostly northeast and some northwest trending subvertical faults, whereas FMI data show that most of the fractures were oriented toward the northwest, with some in the northeast direction. On the basis of the trend of the earthquake locations, faults, and fracture data, the tectonic episodes that created the Central Kansas Uplift and the Nemaha Ridge Humboldt Fault Zone also affected Wellington Field and generated similar structures in that area as well. Observations We found that the long-period, long-duration events (tremor events) occurred when the CO2 injection was being performed between January and June of 2016. Kumar et al. (6) suggested that the filtered waveforms of these signals bear visual similarity to those of tectonic tremors found in geologic settings under natural tec- tonic stresses. These events contain the highest energy in 1 to 5 Hz, which is part of the 1- to 12-Hz frequency band in which tectonic tremors are usually visible. The events are not visible in regional seismic networks in this region, sug- gesting their local origin. By beamforming the tremor signals in 1 to 5 Hz, we showed that they have a low slowness value (<0.3 s/km), indicat- ing their subsurface origin. From the initial catalog of Kumar et al. (6), we selected several events with high signal-to-noise ratios that passed our strict quality tests (fig. S3). Using the wave- form characteristics of these events, we detected additional events not previously recognized during the injection period. Most of the signals had durations ranging from 2 to 5 min. After conducting thorough quality evaluations on these events, we ultimately selected 27 to locate using the beam backprojection technique (24–27). The locations of the signals were deter- mined by assuming the depth in which the supercritical CO2 injections took place. This assumption is necessary because a single array cannot constrain depth of the signal indepen- dently. All of the locations that we obtained were within 5 km of the injection well. The spatial pattern of the locations is roughly aligned along northeast-southwest and northwest- southeast directions. These trends are also shared by the faults of the Central Kansas Uplift and the Nemaha Ridge Humboldt Fault Zone in this region. Schwab et al. used exploration seismic data to show that basement faults did extend to the horizons where the fluids were being injected (14). Schwab et al. also showed the presence of natural open fractures in the target formations, as shown by the FMI data from Fig. 1. Location of the network in relation to earthquake occurrences and other tectonic structures in Kansas. (A) Location of the ZA network in Wellington Field. We analyzed three-component seismic data from December 2015 to July 2016. The area is shown by the blue rectangle in the inset map. (B) Historical and present earthquake occurrences and main structural elements in Kansas. The red highlighted box indicates Sumner County, and the yellow circle indicates Wellington Field. Earthquake magnitudes are represented by the area of the circles, and color indicates the distribution of historical and present-day seismicity. Most of the earthquakes follow the main structural trends. Figure is modified from Hildebrand et al. (53). wells 1-28 and 1-32. The northeast and north- west trend of the source locations of our events was similar to the distribution of regional earth- quakes in the USGS catalog and was also shown by the open fractures in the wells. We observed that the occurrence of the tremor signal and local earthquakes had no temporal correlation between them, even though their spatial trends were along the same northeast and northwest directions. No observable in- crease in the local background seismicity occurred during or after the supercritical CO2 injection. Most earthquakes in the south- central Kansas area stem from oil and gas operations, mainly from the disposal of waste- water (28, 29). The lack of increase in con- temporary earthquakes rates during the CO2 injection may indicate that the injection rate was not high enough to generate additional earthquakes in the region. However, the dis- tribution of tremor signals was limited to the supercritical CO2 injection period, and we did not detect these signals in the month before or after the injection period. The most distinct characteristic of tremor signals is the presence of narrow, more or less Niyogi et al., Science 381, 553–558 (2023) 4 August 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Similarity of the spectrogram characteristics between tremors found in this study and those found in other geologic environments. A spectrogram is a visual representation that shows how the energy in different frequencies changes over time in the data, with the x and y axes representing time and frequency, respectively. Here, we show a spectrogram of volcanic tremors (A) reported by Almendros et al. (54) at Arenal Volcano in Costa Rica, seismic tremors associated with glacial earthquakes (B) at Whilllans Ice Stream in Antarctica described by Winberry et al. (33), and two examples of the tremor events detected in this study [(C) and (D)]. Note the gliding spectral lines in the spectrogram at 1 to 5 Hz. The presence of analogous spectrogram signatures within comparable frequency ranges highlights the similarity in the mechanisms generating these events. evenly spaced bands of frequency that gradu- ally increase, reach a peak, and then decrease (Fig. 2). These time-varying spectral lines are also seen in tremor signals from glacial and vol- canic regions (30, 31) (Fig. 2). The spectral lines are characterized by relatively high amplitude and generally stay within the 1- to 5-Hz range. The peak frequency often aligns with the peak amplitude in the waveform in the time domain. Typically, we observed two spectral lines, al- though three lines were occasionally visible in the spectrogram. A physical model to explain tremors Mathematical models have been proposed to explain the occurrence of harmonic tremors de- tected near volcanoes. Dmitrieva et al. (32), using the general equations given by Hotovec et al. (31), derived and then showed that a delta func- tion with increasing recurrence interval as a function of stress accumulation rate could generate the characteristic harmonic tremor signal seen near volcanoes. Dmitrieva et al.’s model shows that the signal is generated by shear slip induced by fluid in the volcanic sys- tem. In another study, Lipovsky and Dunham (30) showed that tremors detected by seismo- meters in Whillans Ice Plain, West Antarctica, occurred during large-scale sliding events, and this was confirmed by GPS stations. These tremor events are also characterized by gliding spectral lines and can be modeled on the basis of repeating earthquake patches at the ice-bed interface. Our tremor events bear similarities to both volcanic (31) and glacial tremors (30, 33) in terms of the gliding spectral lines. The sig- nals for both the volcanic and glacial tremors are produced by frictional processes. Signals showing discernable narrow bands of varying frequency with time are related to a varying re- currence interval of events, which in turn is related to the changing stressing rate a on the fault plane (32). We used the same mathemati- cal model to show that a pattern of increasing followed by decreasing stressing rate a can ex- plain the occurrence of the tremor signals within the 1 to 5 Hz identified in this study. According to Dmietrieva et al. (32), the spec- trum of N repeating events occurring periodi- cally with a recurrence interval value of T is given by multiplying an individual spectrum event by a spectrum of sequence of N delta functions [as derived in Hotovec et al. (31)]: υ wð Þ e (cid:2)iw M0 4prc3r exp (cid:2) (cid:3) iwr c (cid:2) exp (cid:2) wj jr 2cQ (cid:3) sin (cid:5) (cid:4) wNT (cid:5) (cid:4) 2 wT sin 2 where the variable w is the angular frequency, u(w) is the Fourier transform of particle veloc- ity u(t), M0 is the seismic moment, r is the density of the propagating medium, r is the propagation distance, c is the shear wave speed, and Q is the quality factor. Niyogi et al., Science 381, 553–558 (2023) 4 August 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Analysis of the detected events through examination of nearby earthquake activities and injection program duration. (A) Source locations of tremor events as determined by the beam backprojection method (25, 26). Some tremor locations are aligned with the Central Kansas Uplift (northwest- southeast) and others with the Humboldt Fault Zone (northeast-southwest). Regular earthquakes are aligned with the same structures. (B) Modeling of the tremor signal. Shown is an illustration of a series of earthquakes with temporally varying recurrence interval times as a result of increasing and decreasing rates of stress accumulation. (C) Illustration of multiple spectral lines as a result of the series of earthquakes shown in (B). A quadratic rate of change of events can simulate the pattern seen in the spectrograms of the observed tremor signals. (D) Expanded timeline of the histogram of the tremor events. (E) Occurrence of tremor events (black) plotted in the backdrop of earthquakes (blue) occurring within 50 km of the injection well. (F) Plot of the maximum magnitude earthquakes (green), along with the cumulative number of earthquakes (red). Dmitrieva et al. (32) showed that the spec- tral peaks occur when wT/2 becomes an inte- ger multiple of p. The recurrence interval T is dependent upon stress t and stress accumula- tion rate a as T ∼ dt/a. Considering nearly equal stress drop value for events generated on a fault plane, an increase in stress accu- mulation rate a leads to a decrease in T and vice versa. Any smooth change in the recurrence interval rate is reflected in the spectrograms of these events as spectral lines increasing or decreasing in frequency as a function of time (equidistant signals are shown in Fig. 2, C and D, and fig. S3). Winberry et al. (33) reported the timings of the spectral crests of the observed tremors, and these match the highest rate of glacial earthquake occurrences. This compari- son indicates, as expected, that the occurrences of tremor signals with gliding frequencies are not restricted to volcanic environments alone. We modified some of the parameter values in Dmitrieva et al. (32) for modeling our tremor events because the frequency ranges of volcanic tremors and our events overlap. In our model- ing, we also assumed a linear and quadratic change of recurrence interval value T. We found that multiple spectra with different starting and ending recurrence intervals generate signals that most closely resemble the ones seen in Kansas. As observed in the spectrograms, the Niyogi et al., Science 381, 553–558 (2023) 4 August 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E chief characteristics of the signal, the gliding equidistant spectral lines, can be modeled by multiple subsequent earthquakes with varying interevent times. Results and discussion Even though earthquakes have increased in the northern Oklahoma and southern Kansas region overall due to subsurface fluid injection (13, 34–42), we found no obvious change of background seismicity at the beginning of the CO2 injection program for oil recovery in the Wellington Field region (Fig. 3E and fig. S2). We also saw evidence of natural fractures in those formations from borehole image logs oriented in the same northeast-southwest direc- tion as the Humboldt Fault Zone. Previous stud- ies showed that alterations in pore pressure resulting from fluid injections can trigger both seismic and aseismic activity in fault zones provided that the local stress conditions, hydro- mechanical coupling, and fluid diffusion pro- cesses are conducive to such occurrences (43). Thus, the tremor signals with gliding spectral lines in our study area can provide an expla- nation for why faults are releasing some of the accumulated stress without causing an increase in regular seismicity, i.e., earthquakes. We have determined that the characteristics of the wave- forms and spectra of the signals can be modeled as a series of shear slip events likely induced by fluid injection. On the basis of these find- ings, we infer that these tremor sources are driven by aseismic slip. Previous studies have shown that harmonic tremors are associated with stress state changes caused by the movement of magma and fluids in the subsurface (31, 44). Theoretical models such as those developed by Thomas and Neuberg (45) are valuable for studying source mechanisms because they are based on the observations of harmonic tremors. Therefore, we derived a physical model that is based on Dmitrieva et al. (32) and inferred that a temporally varying stress pattern can generate the tremor signals with gliding spectral lines observed in this area. Aseismic movement of faults can occur aided by increases in pore fluid pressure near the faults caused by fluid injections. Past studies showed that frictional properties can change during fluid injection from rate weakening to rate strength- ening with increasing fluid pressure, which helps to maintain aseismic slip (10). A correlation also exists among earthquakes, aseismic slip, and the pumping of fluids near subsurface faults, as shown in areas such as Delaware Basin in west Texas (46, 47). We relate the lack of increased seismicity in Wellington Field to aseismic fault movements that may have driven the localized tremor sources found in this area. The occurrences of these tremor events have implications for the subsurface movement of injected fluids. Tracking the movement of sub- surface fluids is usually done through simu- lations and modeling (48, 49). Observations of these low-frequency tremor events and their source locations can help us track deformation, fracture location, and movement of injected fluids in the subsurface around the injection well. Determining the locations of induced earthquakes is one of the ways to deduce the extent of subsurface fluid migration from the injection well (13, 50–52). The injection rate of fluids for disposal and enhanced oil recovery purposes is monitored carefully to prevent in- duced earthquakes from increasing in frequency and magnitude. Therefore, observations of such tremor events, their source locations, and their spatiotemporal distribution may provide a use- ful and benign way of tracking the migration of injected fluids. Conclusions We analyzed tremor signals and showed that they originated from the subsurface during a CO2 injection program in Wellington Field in Sumner County, Kansas. Using array analyses, we ruled out surface noises such as freight trains and helicopters as being the origin of these tremors. By locating the events using the beam backprojection method, we were able to show that all of the events occurred in the vicinity of the injection well and that they followed the major structural trends shown by the Humboldt Fault Zone and the Nemaha Uplift. This trend was also evident in the spatial pattern of earthquakes in Kansas over a long time scale. Based on the similarities in their waveforms and spectral characteristics, the closest physical analog of these signals are tremors with gliding spectral lines found in volcanic and glacial settings. These tremor sig- nals are modeled as being generated by a series of small earthquakes with temporally varying recurrence interval times. We infer that these tremors may be driven by aseismic slow slip. Such induced tremor events provide a way to track stress distribution and deformation dur- ing subsurface high-pressure fluid injection. RE FERENCES AND NOTES 1. D. Bame, M. Fehler, Geophys. Res. Lett. 13, 149–152 2. 3. (1986). I. Das, M. D. Zoback, Geophysics 78, KS107 (2013). I. Das, M. D. Zoback, Geophysics 78, KS97–KS108 (2013). 4. H. Hu, A. Li, R. Zavala-Torres, Geophys. Res. Lett. 44, 4814–4819 (2017). 5. P. K. Ghahfarokhi et al., Interpretation (Tulsa) 7, SA1–SA10 (2019). 6. A. Kumar et al., in SEG Technical Program Expanded Abstracts 2019 (Society of Exploration Geophysicists, 2019), pp. 4903–4907. 7. L. Burlini, G. Di Toro, P. Meredith, Geophys. Res. Lett. 36, L08305 (2009). 8. T. Yamashita, Geophys. J. Int. 193, 375–393 (2013). 9. M. Kano, A. Kato, R. Ando, K. Obara, Sci. Rep. 8, 3655 (2018). 12. H. Yu, R. M. Harrington, H. Kao, Y. Liu, B. Wang, Nat. Commun. 12, 6862 (2021). 13. S. L. Peterie, R. D. Miller, J. W. Intfen, J. B. Gonzales, Geophys. Res. Lett. 45, 1395–1401 (2018). 14. D. R. Schwab, T. S. Bidgoli, M. H. Taylor, J. Geophys. Res. Solid Earth 122, 10,132–10,154 (2017). 15. J. M. Gregg, K. L. Shelton, in The Great American Carbonate Bank: The Geology and Economic Resources of the Cambrian– Ordivician Sauk Megasequence of Laurentia, J. Derby, R. Fritz, S. Longacre, W. Morgan, C. Sternbach, Eds. (AAPG, 2012); https://doi.org/10.1306/13331493M981487. 16. R. C. Gutschick, C. A. Sandberg, “Mississippian continental margins of the conterminous United States” (Society of Economic Paleontologists and Mineralogists, 1983), SEPM Special Publication No. 33, pp. 79–96; https://archives. datapages.com/data/sepm_sp/SP33/Mississippian_ Continental_Margins.htm. 17. S. Mazzullo, B. W. Wilhite, B. T. Morris, D. Boardman, “Syndepositional tectonism and its effects on Mississippian (Kinderhookian to Osagean) lithostratigraphic architecture: Part 2 – Subsurface occurrences in the midcontinent USA. AAPG Search and Discovery Article 30208,” paper presented at AAPG Mid-Continent Section meeting, Oklahoma City, Oklahoma, 1–4 October 2011 (2011), Search and Discovery Article #30208; https://www.searchanddiscovery.com/ documents/2011/30208mazzullo/ndx_mazzullo.pdf. 18. D. F. Merriam, The Geologic History of Kansas (Kansas Geological Survey, 1963), vol. 162. 19. A. Layzell et al., Curr. Res. Earth Sci. 263, 1–8 (2017). 20. W. Lee, Stratigraphy and Structural Development of the Salina Basin Area (Kansas Geological Survey, 1956), vol. 121. 21. S. M. Rogers, AAPG Bull. 85, 115 (2001). 22. B. W. Wilhite, S. Mazzullo, B. T. Morris, D. Boardman, “Syndepositional tectonism and its effects on Mississippian (Kinderhookian to Osagean) lithostratigraphic architecture: Part 1 – Based on exposures in the midcontinent USA,” paper presented at AAPG Mid-Continent Section meeting, Oklahoma City, Oklahoma, 1–4 October 2011 (2011), Search and Discovery Article #30207; https://www.searchanddiscovery. com/documents/2011/30207wilhite/ndx_wilhite.pdf. 23. D. E. Zeller, The Stratigraphic Succession in Kansas (Kansas Geological Survey, 1968), vol. 189. 24. A. Ghosh, J. E. Vidale, J. R. Sweet, K. C. Creager, A. G. Wech, Geophys. Res. Lett. 36, L17316 (2009). 25. A. Ghosh, J. E. Vidale, K. C. Creager, J. Geophys. Res. Solid Earth 117, B10301 (2012). 26. A. A. Hutchison, A. Ghosh, Bull. Seismol. Soc. Am. 107, 1985–1993 (2017). 27. B. Li, A. Ghosh, Geophys. Res. Lett. 44, 5427–5435 (2017). 28. J. L. Rubinstein, W. L. Ellsworth, S. L. Dougherty, Bull. Seismol. Soc. Am. 108, 674–689 (2018). 29. R. M. Pollyea, M. C. Chapman, R. S. Jayne, H. Wu, Nat. Commun. 10, 3077 (2019). 30. B. Lipovsky, E. Dunham, Cryosphere 10, 385–399 (2016). 31. A. J. Hotovec, S. G. Prejean, J. E. Vidale, J. Gomberg, J. Volcanol. Geotherm. Res. 259, 89–99 (2013). 32. K. Dmitrieva, A. J. Hotovec-Ellis, S. Prejean, E. M. Dunham, Nat. Geosci. 6, 652–656 (2013). 33. J. P. Winberry, S. Anandakrishnan, D. A. Wiens, R. B. Alley, Geophys. Res. Lett. 40, 312–315 (2013). 34. K. M. Keranen, H. M. Savage, G. A. Abers, E. S. Cochran, Geology 41, 699–702 (2013). 35. A. L. Llenos, A. J. Michael, Bull. Seismol. Soc. Am. 103, 2850–2861 (2013). 36. M. Schoenball, W. L. Ellsworth, J. Geophys. Res. Solid Earth 122, 10,189–10,206 (2017). 37. L. Johann, S. A. Shapiro, C. Dinske, Sci. Rep. 8, 11505 (2018). 38. J. H. Norbeck, J. L. Rubinstein, Geophys. Res. Lett. 45, 2963–2975 (2018). 39. G. Zhai, M. Shirzaei, M. Manga, X. Chen, Proc. Natl. Acad. Sci. U.S.A. 116, 16228–16233 (2019). 40. C. Joubert, R. Sohrabi, J. L. Rubinstein, G. Jansen, S. A. Miller, Geophys. Res. Lett. 47, e2020GL088326 (2020). 41. A. Amemoutou, P. Martínez-Garzón, G. Kwiatek, J. L. Rubinstein, M. Bohnhoff, J. Geophys. Res. Solid Earth 126, e2020JB021625 (2021). 10. F. Cappa, M. M. Scuderi, C. Collettini, Y. Guglielmi, J.-P. Avouac, 42. A. Verdecchia, E. S. Cochran, R. M. Harrington, J. Geophys. Sci. Adv. 5, eaau4065 (2019). Res. Solid Earth 126, e2020JB020384 (2021). 11. L. De Barros, Y. Guglielmi, D. Rivet, F. Cappa, L. Duboeuf, 43. F. H. Cornet, Geomechanics for Energy and the Environment 5, C. R. Geosci. 350, 464–475 (2018). 42–54 (2016). Niyogi et al., Science 381, 553–558 (2023) 4 August 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 44. Y. Yukutake et al., Earth Planets Space 69, 164 (2017). 45. M. E. Thomas, J. Neuberg, Geology 40, 351–354 (2012). 46. S. P. Hicks, S. Goes, A. C. Whittaker, P. J. Stafford, J. Geophys. Res. Solid Earth 126, e2021JB022768 (2021). 47. K. S. Pepin, W. L. Ellsworth, Y. Sheng, H. A. Zebker, J. Geophys. Res. Solid Earth 127, e2021JB023157 (2022). 48. A. K. Schwartzkopff, A. Sainoki, D. Elsworth, Rock Mech. Rock Eng. 54, 1027–1053 (2020). 49. A. K. Schwartzkopff, A. Sainoki, D. Elsworth, Int. J. Rock Mech. Min. Sci. 147, 104871 (2021). 50. W. L. Ellsworth, Science 341, 1225942 (2013). 51. K. M. Keranen, M. Weingarten, G. A. Abers, B. A. Bekins, S. Ge, Science 345, 448–451 (2014). 52. F. R. Walsh 3rd, M. D. Zoback, Sci. Adv. 1, e1500195 (2015). 53. G. M. Hildebrand, D. W. Steeples, R. W. Knapp, R. D. Miller, B. C. Bennett, Seismol. Res. Lett. 59, 159–163 (1988). 54. J. Almendros, R. Abella, M. M. Mora, P. Lesage, J. Geophys. Res. Solid Earth 119, 5536–5559 (2014). 55. S. Niyogi, A. Ghosh, A. Kumar, R. W. Hammack, Data for: Tremor signals during fluid injection are generated by fault slip, Zenodo (2023); https://doi.org/10.5281/zenodo.8045320. ACKN OWLED GMEN TS We thank the two anonymous reviewers for their dedication and time in going through the manuscript, highlighting the necessary changes, and recommending improvements to the text. Their suggestions have substantially improved this manuscript. Funding: This work was supported by the University of California Riverside. Author contributions: S.N. analyzed the data, performed the experiments, and contributed to the writing of the manuscript. A.G. conceived the study, planned the experiments, coordinated the project, and assisted in analyzing the data, interpreting the results, and writing the manuscript. A.K. and R.H. provided the waveform data and station metadata and assisted in interpreting the results. All authors discussed the manuscript and contributed to the science presented in the article. Competing interests: The authors declare no competing interests. Data and materials availability: Waveform data and codes can be accessed at Zenodo (55). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh1331 Materials and Methods Figs. S1 to S6 References (56–64) Submitted 13 February 2023; accepted 23 June 2023 10.1126/science.adh1331 Niyogi et al., Science 381, 553–558 (2023) 4 August 2023 6 of 6
10.1126_science.adh7670
RES EARCH FERROELECTRICS Atomic-scale polarization switching in wurtzite ferroelectrics Sebastian Calderon V1, John Hayden2, Steven M. Baksa2, William Tzou1, Susan Trolier-McKinstry2, Ismaila Dabo2, Jon-Paul Maria2, Elizabeth C. Dickey1* Ferroelectric wurtzites have the potential to revolutionize modern microelectronics because they are easily integrated with multiple mainstream semiconductor platforms. However, the electric fields required to reverse their polarization direction and unlock electronic and optical functions need substantial reduction for operational compatibility with complementary metal-oxide semiconductor (CMOS) electronics. To understand this process, we observed and quantified real-time polarization switching of a representative ferroelectric wurtzite (Al0.94B0.06N) at the atomic scale with scanning transmission electron microscopy. The analysis revealed a polarization reversal model in which puckered aluminum/boron nitride rings in the wurtzite basal planes gradually flatten and adopt a transient nonpolar geometry. Independent first-principles simulations reveal the details and energetics of the reversal process through an antipolar phase. This model and local mechanistic understanding are a critical initial step for property engineering efforts in this emerging material class. B etween 2019 and 2021, researchers world- wide demonstrated unexpected ferro- electricity in solid solutions in the AlN-ScN (1), AlN-BN (2), GaN-ScN (3), and ZnO- MgO (4) composition families. These re- ports disrupted the more than 100-year-old perspective that wurtzite crystals are pyro- electric and piezoelectric but cannot be ferro- electric because the polarization cannot be switched with an electric field. The important scientific and technological consequences that accompanied this understanding were that (i) we have a previously unknown class of ferro- electrics that violates the long-standing rela- tionship linking polarization and ferroelectric transition temperature (5)—that is, they are likely not soft-mode ferroelectrics; (ii) the new physical mechanisms that enable ferroelectricity 1Department of Materials Science and Engineering, Carnegie Mellon University, Pittsburgh, PA 15213, USA. 2The Pennsylvania State University, Department of Materials Science and Engineering and Materials Research Institute, University Park, PA 16802, USA. *Corresponding author. Email: ecdickey@cmu.edu will almost certainly impart different manifes- tations of physical scaling trends and differ- ent property dependencies on temperature, pressure, and time; and (iii) these materials can be processed at or near room temperature with a robust property response, and in some cases (for example, Al1−xBxN) they consist ex- clusively of elements that are already common and essential in the silicon complementary metal-oxide semiconductor (CMOS) front end. This emerging ability to directly integrate a strongly hysteretic nonlinear dielectric is trans- formational for future logic, memory, high- power, acoustic, or electro-optic devices (6–8). These ferroelectrics, however, bring their own challenges, the most notable being that the margin between the coercive field and break- down field is uncomfortably small at room tem- perature. Consequently, devices are operated in close proximity to their failure thresholds, which invariably erodes electrical endurance and can induce degradation pathways. Lowering the energetic switching barriers is a principal goal necessary to realize the full po- tential of ferroelectric wurtzites in electrical, optical, and acoustic devices. Empirically, we know that epitaxial tensile strain and an in- creased dopant concentration can reduce coer- cive fields in Al1−xScxN (9, 10) and Al1−xBxN (2), presumably by reducing the energy barriers that regulate the polarity switch, and that in- creasing temperature also reduces the coercive field by thermally activating the nucleation and switching processes (11). These trends are based on macroscopic observations—that is, on measuring millimeter-scale capacitors—and although the trends demonstrate possibilities to reduce the coercive field, they do not pro- vide local mechanistic insight. Understanding these trends and phenomena at global and local-length scales is essential so that the property response of wurtzite ferro- electrics can be further developed for practical use. We addressed this challenge by measur- ing in situ the polarization reversal process at the atomic scale. Our experimental approach takes advantage of local electrostatic sample charging occurring during the illumination of dielectric materials in the electron microscope, which can induce the necessarily large local electric fields (12, 13) for polarization switch- ing in ferroelectrics (14, 15). We used scanning transmission electron microscopy (STEM) in differentiated differential phase contrast mode (dDPC) to observe in situ polarization inversion in thin films of the composition Al0.94B0.06N [(Al,B)N]. Materials in this composition space have remanent polarization (Pr) values exceed- ing 125 mC cm–2 while maintaining bandgaps above 5.8 eV; this is a substantial advantage relative to other wurtzite-based ferroelectrics, albeit the coercive fields are also larger (2). Our experimental analysis documents polarization switching with quantitative information at the atomic scale that is coinformed with first- principles calculations of the same compo- sition. The first-principles simulation analyses provide an atomic-scale understanding of the switching pathways and energetics in these emerging ferroelectric materials. Covalidation Fig. 1. Structure of (Al,B)N. (A) dDPC-STEM images for (Al,B)N film. (B) Fast Fourier transform calculated from the (Al,B)N region of image in (A). (C) dDPC-STEM image of an (Al,B)N grain oriented along the [110] direction, where the location of the atomic columns has been highlighted with blue (N) and gray (Al) circles after Gaussian fitting. “P” with arrow shows the direction of polarization. (D) dDPC image overlapped with a vector map indicating the calculated polarization at each unit cell. Calderon et al., Science 380, 1034–1038 (2023) 9 June 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E of STEM imaging and density functional the- ory support a switching mechanism in which a transient antipolar structure mediates the switching process; the transient structure is similar to long-standing models for inversion domain boundaries in III-nitrides. Effect of boron and wake-up process on the structure We show a dDPC-STEM image of an (Al,B)N film grown by means of reactive pulsed dc sput- tering on a W electrode, showing a columnar grain structure with a nominal epitaxial relation of (Al,B)N <001> // W <110> // growth direction (Fig. 1). The W electrode is faceted, and we observed small variations in the orientation of (Al,B)N grains, leading to the low density of dislocations that are observed in the image. We detected the out-of-plane crystalline mosaicity in the rocking curves of the 002 reflection of the (Al,B)N films, with full width at half-maximum (FWHM) values of 1.44° (fig. S1). From the STEM images, we made structural polarization measurements to compare with macroscopic measurements. To quantify the spontaneous polarization, we acquired dDPC images along the <110> zone axis (Fig. 1C), where the growth direction points from the bottom to the top of the image along the 00(cid:1)1 (cid:2) direction of (Al,B)N. We determined the polar- ization orientation from the images by estab- ½ lishing the nitrogen tetrahedral orientation (Fig. 1C, inset). This orientation is polarization down—that is, the positive end of the perma- nent dipole points to the substrate, which is commonly referred to as an “N-polar” growth direction. Among the observed regions, the N-polar orientation is uniform with no evi- dence of Al-polar orientations, which is consist- ent with prior studies that show uniformly N-polar-oriented films (16). The atomic-column positions of both Al(B) and N were determined with a Gaussian fitting process on the dDPC images, which allows precise location of the atomic-column positions and thus an accurate quantification of the polarization, as previously reported (17). The polarization was quantified by using the Born effective charges calculated from density-functional perturbation theory (18). We show a vector-map representation of the calculated polarization per unit cell for an as-deposited (Al,B)N film (Fig. 1D), which we discuss in comparison with other samples. We constructed a summary comparison of the calculated polarization per unit cell for as-deposited pure AlN film (Fig. 2A), an as- deposited (Al,B)N film (Fig. 2B), and a field- cycled (Al,B)N film left in the opposite polarity from the as-deposited film (Fig. 2C). The sputtered AlN film (Fig. 2A) exhibits an average polarization magnitude of 1.13 ± 0.07 C m−2. The standard deviation is close to the measure- ment uncertainty arising from the atomic- column fitting process. The average measured polarization magnitude is 6.6% less than that of the ideal AlN structural model (1.21 C m−2). The as-deposited (Al,B)N film (Fig. 2B) exhib- its an average spontaneous polarization of 1.30 C m−2, 13% larger than the undoped film, which is in agreement with the trend pre- dicted by first principles (2). The (Al,B)N film exhibits peak-to-peak variations of 1.13 to 1.36 C m−2 and a standard deviation of 0.10 C m−2 within the measured area. These variations, which exceed the measurement error, are consistent with the chemical and structural inhomogeneities associated with the 6% B substitution—for example, FWHM x-ray peak widths of ~1° in w and 2q circles and a large radial misfit of B for Al. Electric field–induced polarization switching from the as-deposited unipolar state can be highly frequency dependent. For kilohertz pulses just above the coercive field, full reversal per cycle may take many tens of cycles. For slow cycling (for example, 0.1 Hz), one sweep can be sufficient. This “wake-up” process that precedes fast polarization reversal is ascribed to low initial domain-wall density and mobility (16). To better understand this wake-up pheno- menon, we completed the same STEM polar- ization analysis on a field-cycled sample and compared those results against bulk electrical Fig. 2. Polarization evolution. (A to C) dDPC images for (A) pure AlN, (B) (Al,B)N as-deposited, and (C) field-cycled (Al,B)N. (D) Schematic of the structure modification after boron incorporation and field cycling. dDPC-STEM images showing the full field of view are presented in fig. S2. Calderon et al., Science 380, 1034–1038 (2023) 9 June 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. In situ polarization switching. (A) dDPC-STEM frames at 0, 209, 266, and 361 s for in situ polarization switching. (B) AlN dumbbell-angle evolution as a function of time for the corresponding frames in (A). The Roman numerals in (A) and (B) indicate the different regions in the images going from the top of the frame (I) to the bottom (IV). The rows of AlN dumbbells in the images are numbered from 1 to 8 for reference between (A) and (B). (C) Dumbbell-angle evolution as a function of time for one unit cell. (Inset) The schematics of the dumbbell angle and the dDPC images at specific times. measurements. For a fully woken-up sample, hysteresis loops yield Pr of 1.38 ± 0.01 C m−2, whereas dDPC-STEM imaging yields 1.40 ± 0.14 C m−2. Furthermore, the STEM-based polarization measurements allow us to com- pare the intrinsic spontaneous polarization in the as-deposited and field-cycled states (Fig. 2, B and C). The similarity of these STEM- quantified polarization magnitudes (average values within ~7%) suggests that the wake-up process does not substantially alter the bulk material structure. To further explore this possibility, we ana- lyzed the average local structure evolution using two-dimensional vector pair correlation functions (vPCF) calculated from dDPC-STEM images that can be found in the supplemen- tary materials (fig. S3) (19). The relevant param- eter extracted from vPCF is Dr (Fig. 2D), which describes the distance between the nearest (Al,B) basal-plane layers and N layers, also referred to as the “puckering” of the wurtzite basal planes. Whereas the average Dr increases 10% on alloying AlN with 6% B, the Dr in as- deposited and field-cycled (Al,B)N states are the same within one standard deviation of the measurement distribution, supporting the con- clusion of no major structural phase transitions during the wake-up process. We therefore infer that the wake-up phenomenon is more likely associated with interfacial or nucleation site–based mechanisms than bulk crystal- structure changes. In situ switching observations Electron microscopy can visualize ferroelectric polarization reversal at length scales that reveal local atomic and domain structures (15, 20). In these cases, preparation of extremely thin parallel-plate capacitor structures that can be field cycled either in situ or by imple- menting a nanoprobe that applies a local bias is common (20). Wurtzite ferroelectrics are difficult to analyze with these methods because of the large coercive fields that invariably accelerate surface migration of atomic species across the sample surface and create short- circuit pathways. Alternatively, prior reports demonstrate that the charge accumulation from beam-solid interactions in electron and ion microscopes can create internal fields large enough to reverse ferroelectric polarization (14, 15). By using extended beam exposures during STEM image acquisition, polarization reorientation can be achieved from the large lateral electric fields that emerge from the beam-solid interactions that produce positive sample charging, as described in detail by Cazaux (12). For very thin samples, incident electrons do not contribute substantially to charge accumulation because they are fully transmitted. However, some electrons are scattered inelastically, promoting ionization and secondary and Auger electron emission and a net positive charge accumulation for electrically insulating materials. This super- imposes additional built-in electric fields, which can reach the megavolt-per-centimeter range needed to switch (Al,B)N. Our primary data- set is an in situ domain switching movie collected on an (Al,B)N film (movie S1). We summarize the in situ data with a set of frames spanning the roughly 7-min experiment (Fig. 3). The images are indexed from top to bottom with row numbers and divided into four regions (I, II, III, and IV), with region I closest to vac- uum (the top of the sample), where the charge- induced field is expected to be largest in the vertical direction (and thus capable of inducing switching), and region IV closest to the bottom electrode, which provides a physical path to Calderon et al., Science 380, 1034–1038 (2023) 9 June 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Nudged-elastic-band simulation of polarization reversal pathways. (A) Nudged-elastic-band simulation of polarization reversal pathways for AlN and Al15/16B1/16N. (Insets) The structural models at specific stages of the simulation. (B) Structural model of the nonpolar transient state calculated for Al15/16B1/16N viewed along different projections. (C) Atomic models, STEM image simulations, and experimental images for the N-polar, nonpolar, and Al-polar states. microscope ground. This distinction is impor- tant because the electric field in the experiment will necessarily be spatially heterogeneous in magnitude and evolve over time; nonetheless, it serves as an important mechanism to in- duce switching in these large-coercive-field samples. Our STEM images are overlayed with red, white, and blue lines, which indicate quan- titatively (from image analysis) the dumbbell angles associated with the bond puckering in the wurtzite basal plane; red is positive, blue is negative, and white is 0° on average. We show four graphs that quantify the average dumb- bell angles for regions I through IV (Fig. 3B), each one representing a unit-cell-thick layer. We note four observations: (i) In region I, the transition between red and blue is the most clear and uniform and occurs first; this observation is consistent with expectations for maximum charging fields in this zone. (ii) As one progresses from the top zones to the bot- tom zones, the dumbbell-angle transition is less uniform, and the switching process takes longer; this is again consistent with the antic- ipated electric-field evolution and gradient direction. (iii) Region IV and below do not completely switch, which is attributed to dimin- ishing electric field toward the bottom of the sample. (iv) The regions of the sample that switch do so through a transient state in which the integrated intensities of all associated scatterers yield an average dumbbell angle of 0° (Fig. 3C). We also note the 209-s panel in Figure 3A, and in particular, the yellow highlighted rectan- gular region where the largest electric field is expected. The dumbbell-angle rotation, and thus switching, initiates first at the top of this rectangle and progresses laterally and verti- cally with time; the structure in this region is identified as a local domain nucleation site with attendant domain walls. Last, we show the average AlN dumbbell- angle progression for one unit cell in the movie close to the location of maximum anticipated electric field (Fig. 3C). The angle evolution from positive to negative values, and thus polariza- tion down to up, is clearly visible. Figure 3C also contains inset images of the initial N-polar (polarization-down) state, the final Al-polar (polarization-up) state, and the transient inter- mediate state of this region of the sample. To further understand the switching path- way and identify the transient state observed experimentally, we carried out first-principles calculations using the nudged-elastic-band method for determining the minimum energy pathway for polarization reversal in both AlN and Al15/16B1/16N, which closely matches the experimental composition. After structural optimization, we calculated the polarization reversal pathways (Fig. 4A). The results high- light two distinct mechanisms, in which AlN switches coherently in one step at 0.523 eV/ formula unit (f.u.), whereas Al15/16B1/16N switches sequentially in nine steps at 0.200 eV/f.u. The coherent polarization reversal suggests that pure AlN switches through a nonpolar hexagonal (h)-BN–like structure, whereas in Al15/16B1/16N, we observed an average non- polar supercell in the middle of the switching process (Fig. 4A, insets). This metastable state is not a h-BN–like structure but consists of anti- polar arrangements of wurtzite motifs when viewed along the [100] direction, as shown in our atomic models (Fig. 4B). A more convoluted structure is observed when projecting the structure along the [110] or [010] orientation (Fig. 4B), where N-polar and Al-polar unit cells are overlapped in this projection, which agrees qualitatively well with the transient structure observed during the in situ switch- ing experiments. Both in situ data and first- principles calculation independently show that the anion and cation sublattice contribute com- parable distortions along the switching path. To correlate the experimental observations and the first-principles calculations, we carried out dPDC-STEM image simulations using the atomic models of the Al15/16B1/16N, at the initial (N-polar), middle (nonpolar), and final (Al-polar) step in the simulation (Fig. 4C). The simulated images correlate well with the experimental STEM images at different stages during the switching process, corroborating the nonpolar state predicted with the nudged-elastic-band simulations. Similar projected local structures have been observed with TEM and STEM images in polarity inversion domain boundaries (IDBs) in other wurtzite structures (21–23), and the four- and eightfold rings of bonds that were found in simulations of IDBs in GaN (24) are present in this metastable Al15/16B1/16N structure (fig. S4). The low-energy IDB in GaN exhibits no interface electronic gap states because the Ga and N atoms can maintain similar local coordination as the parent wurtz- ite phase. These results provide insight into the induced ferroelectricity of wurtzite structures such as AlN, where the incorporation of alloying ele- ments, such as B, in the parent wurtzite struc- ture, results in substantial local-bonding and local-structural distortions, as has been pointed out in the (Al,Sc)N system (25). We demonstra- ted that this disorder provides low-energy path- ways to nucleate the switching process and that the switching is mediated by a metastable transient nonpolar state. This nonpolar meta- stable state maintains a local structure similar Calderon et al., Science 380, 1034–1038 (2023) 9 June 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E to the parent wurtzite structure and analogous to the local structure evidenced in IDBs in numerous nonferroelectric wurtzites. The chemical and structural disorder in the (Al,B)N alloys leads to a relatively flat energy land- scape that facilitates switching. Our work provides detailed atomic-scale structure analysis of ferroelectrically active wurtzite-structured (Al,B)N thin films, reveal- ing that B doping leads to a larger spontane- ous polarization relative to pure AlN sputtered films. The study also unambiguously shows the domain nucleation and switching pathway for these emerging ferroelectric materials. The alloying disorder provides a complex but sub- stantially lower-energy-barrier switching pathway to a transient nonpolar state, which engenders ferroelectricity in the class of technologically important polar materials. Such wurtzite- structured materials can be coprocessed with mainstream semiconductors, providing path- ways toward CMOS-compatible integrated fer- roelectrics, piezoelectrics, and electro-optics. Our insights into the atomic-scale processes of domain switching provide guidance to ongoing and future investigations of wurtzite-structured thin films that can lead to targeted property performance. RE FE RENCES AND N OT ES 1. S. Fichtner, N. Wolff, F. Lofink, L. Kienle, B. Wagner, J. Appl. Phys. 125, 114103 (2019). J. Hayden et al., Phys. Rev. Mater. 5, 044412 (2021). 2. 3. D. Wang, P. Wang, B. Wang, Z. Mi, Appl. Phys. Lett. 119, 111902 (2021). 4. K. Ferri et al., J. Appl. Phys. 130, 044101 (2021). 5. S. C. Abrahams, S. K. Kurtz, P. B. Jamieson, Phys. Rev. 172, 551–553 (1968). 6. A. Sebastian, M. Le Gallo, R. Khaddam-Aljameh, E. Eleftheriou, Nat. Nanotechnol. 15, 529–544 (2020). 7. N. Setter et al., J. Appl. Phys. 100, 051606 (2006). 8. A. Fernandez et al., Adv. Mater. 34, e2108841 (2022). 9. M. Noor-A-Alam, O. Z Olszewski, M. Nolan, ACS Appl. Mater. Interfaces 11, 20482–20490 (2019). 10. S. Clima et al., Appl. Phys. Lett. 119, 172905 (2021). 11. W. Zhu et al., Appl. Phys. Lett. 119, 062901 (2021). 12. J. Cazaux, Ultramicroscopy 60, 411–425 (1995). 13. R. F. Egerton, P. Li, M. Malac, Micron 35, 399–409 (2004). 14. A. C. G. Nutt, V. Gopalan, M. C. Gupta, Appl. Phys. Lett. 60, 2828–2830 (1992). 15. J. L. Hart et al., Phys. Rev. B 94, 174104 (2016). 16. W. Zhu et al., Adv. Electron. Mater. 8, 2100931 (2022). 17. S. Calderon, S. D. Funni, E. C. Dickey, Microsc. Microanal. 28, 1–12 (2022). 18. Materials and methods are available as supplementary materials. 19. S. D. Funni et al., APL Mater. 9, 091110 (2021). 20. L. Li, L. Xie, X. Pan, Rep. Prog. Phys. 82, 126502 (2019). 21. T. Akiyma et al., Jpn. J. Appl. Phys. 58 (SC), SCCB30 (2019). 22. N. Stolyarchuk et al., Sci. Rep. 8, 14111 (2018). 23. L. T. Romano, J. E. Northrup, M. A. O’Keefe, Appl. Phys. Lett. 69, 2394–2396 (1996). 24. J. E. Northrup, J. Neugebauer, L. T. Romano, Phys. Rev. Lett. 77, 103–106 (1996). 25. K. Yazawa, J. S. Mangum, P. Gorai, G. L. Brennecka, A. Zakutayev, J. Mater. Chem. C 10, 17557–17566 (2022). 26. S. Calderon et al., Atomic-scale polarization switching in wurtzite ferroelectrics - dDPC data, Zenodo (2023); https://doi.org/10.5281/zenodo.7916420. 27. S. Calderon et al., Atomic-scale polarization switching in wurtzite ferroelectrics - DFT data, Zenodo (2023); https://doi.org/10.5281/zenodo.7916565. AC KNOWLED GME NTS Funding: This material is based on work supported by the Center for 3D Ferroelectric Microelectronics (3DFeM), an Energy Frontier Research Center funded by the US Department of Energy, Office of Science, Office of Basic Energy Sciences Energy Frontier Research Centers program under award DE-SC0021118. E.C.D. and S.C. acknowledge the use of the Materials Characterization Facility at Carnegie Mellon University supported by grant MCF-677785. Author contributions: Conceptualization: S.C.V. and E.C.D. Material production: J.H. and J.-P.M. Electron microscope characterization: S.C.V, W.T., and E.C.D. First-principles calculations: S.M.B. and I.D. Data interpretation: S.C.V., J.H., S.M.B., E.C.D., J.-P.M., I.D., and S.T.-M. Funding acquisition: E.C.D, J.-P.M., I.D., and S.T.-M. Supervision: E.C.D., J.-P.M., and I.D. Writing – original draft: S.C.V. Writing – review and editing: All authors reviewed and edited the draft. Competing interests: The authors declare that they have no competing interests. Data and materials availability: TEM data generated in this study are available at Zenodo (26). First-principles data and models that support the findings of this study are available at Zenodo (27). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https:// www.science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh7670 Materials and Methods Figs. S1 to S4 Movie S1 References (28–44) Submitted 13 March 2023; accepted 10 May 2023 10.1126/science.adh7670 Calderon et al., Science 380, 1034–1038 (2023) 9 June 2023 5 of 5
10.1126_science.adh5339
RES EARCH ENZYMOLOGY Trivalent rare earth metal cofactors confer rapid NP-DNA polymerase activity Victor S. Lelyveld1,2, Ziyuan Fang1,2,3, Jack W. Szostak1,2,3,4,5,6* A DNA polymerase with a single mutation and a divalent calcium cofactor catalyzes the synthesis of unnatural N3′→P5′ phosphoramidate (NP) bonds to form NP-DNA. However, this template-directed phosphoryl transfer activity remains orders of magnitude slower than native phosphodiester synthesis. Here, we used time-resolved x-ray crystallography to show that NP-DNA synthesis proceeds with a single detectable calcium ion in the active site. Using insights from isotopic and elemental effects, we propose that one-metal-ion electrophilic substrate activation is inferior to the native two-metal-ion mechanism. We found that this deficiency in divalent activation could be ameliorated by trivalent rare earth and post–transition metal cations, substantially enhancing NP-DNA synthesis. Scandium(III), in particular, confers highly specific NP activity with kinetics enhanced by more than 100-fold over calcium(II), yielding NP-DNA strands up to 100 nucleotides in length. T he principal chemistry at the core of RNA and DNA metabolism is phospho- diester synthesis. Polymerases generate nascent strands of genetic material by stepwise phosphoryl transfer of nucleo- tides to primer termini, yielding O3′→P5′ phos- phodiester linkages. This template-directed process is conserved across all known biology. Nucleotide 3′ substitutions have therefore been widely regarded as chain terminating for polymerase activity, forming the basis for a class of nucleoside analog drugs (1, 2). In recent work, we demonstrated the direct enzymatic synthesis of an unnatural linkage by substitution of the 3′-OH nucleophile with a 3′-amine, extending the chemistry amenable to polymerase catalysis (1). We reported that 3′-NH2 primer extension can be catalyzed by a modified DNA polymerase, yielding N3′→P5′ phosphoramidate DNA (NP-DNA; Fig. 1, A and B). The large fragment of DNA polymerase I (BF), cloned from the thermophilic soil bacte- rium Geobacillus stearothermophilus (Bst), ac- quires nontrivial levels of N3′→P5′ polymerase activity through two unexpectedly minor sub- stitutions: a single active-site mutation (F710Y) and substitution of its divalent Mg2+ cofactors with Ca2+ (1). However, this level of activity remained around four orders of magnitude slower than native phosphodiester synthesis. Although NP synthesis was detectable in the presence of several divalent alkaline earth metal ions, the pattern of metal cofactor activ- ity was distinct from that found in native phos- 1Center for Computational and Integrative Biology, Department of Molecular Biology, Massachusetts General Hospital, Boston, MA 02114, USA. 2Department of Genetics, Harvard Medical School, Boston, MA 02115, USA. 3Department of Chemistry, University of Chicago, Chicago, IL 60637, USA. 4Howard Hughes Medical Institute, The University of Chicago, Chicago, IL 60637, USA. 5Howard Hughes Medical Institute, Massachusetts General Hospital, Boston, MA 02114, USA. 6Department of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138, USA. *Corresponding author. Email: jwszostak@uchicago.edu phodiester activity (3). Crystal structures of the ground-state NP reaction complex showed a single metal ion in the active site (1), suggest- ing a plausible distinction between the NP mechanism and the classical two-metal-ion mechanism in native phosphodiester activity (4). Whether NP catalysis does in fact rely on a distinct mechanism with a nonclassical co- factor configuration remains to be established, as do the critical factors limiting the kinetics of NP synthesis. Linearly correlated density dynamics in reacting BF crystals during NP synthesis To understand the mechanistic barriers to rapid enzymatic NP-DNA synthesis, we first sought to observe NP bond formation through crystal- lography. In previous x-ray crystal structures of the fully assembled NP polymerase pre- insertion complex, we modeled a single divalent metal ion in the reaction center, occupying a site distal to the primer-terminal nucleophile (1). This site is equivalent to the so-called “B site” in the classical two-metal-ion reaction center and includes inner sphere ligands from the substrate triphosphate moiety, the side chains of Asp830 and Asp653, and the backbone amide of Tyr654. No evidence for an “A-site” metal proximal to the primer 3′-amino terminus was observed in these earlier prechemistry model structures, a finding consistent with the expectation that a neutral amine is the nucleophile in NP synthe- sis. However, earlier structures did not fully recapitulate the active NP reaction center or the product state (1). The A-site metal ion can also be poorly ordered even in structures of the native polymerase reaction complex. Nakamura et al. reported that accumulation of an additional metal ion in a “C site” of poly- merase h, a polymerase Y family member, occurs in a manner linearly correlated with bond formation, suggesting that the product state’s postchemistry metal configuration may be distinct from that observed in the prechem- istry reaction complex (5). This observation was subsequently extended to polymerase X family members (6, 7). To add evidence that a single divalent metal ion cofactor catalyzes unnatural NP synthesis in BF, a polymerase A family member, we performed a single-nucleotide primer extension reaction in intact crystals (Fig. 1, B to D). We crystallized active substrate-bound BF polymerase F710Y/D598A with a 3′–amino- terminal DNA primer and DNA template in the presence of Ca2+ and 2′-deoxyguanosine 5′- triphosphate (dGTP). Reactions were then ini- tiated in the intact crystals by a pH shift from an acidic mother liquor (pH < 6) to a basic soaking liquor at pH 8.8, and reacting crystals were quenched at various times by flash freez- ing (Fig. 1, C and D). Quantitative analysis of electron density dynamics is complicated by crystal-to-crystal variance in diffraction datasets. We therefore incorporated datasets arising from up to four crystals at each of six time points after ini- tiation. By scaling reflections from 19 total crys- tals (resolution 2.0 to 2.7 Å) across the time course (0 to 24 hours), we could estimate voxel- wise first-order rate constants for changes in real-space electron density across the asym- metric unit and particularly at the reaction center (Fig. 1D, bottom left panel). Early time points show negligible differences in ordered density between the ground-state complex crys- tallized under acidic conditions versus the com- plex after 1 hour of soaking under reaction conditions, suggesting that the reactant state and ground state cannot be clearly distin- guished in this resolution range. By following the reaction at subsequent times, we observed monotonic accumulation of nascent NP bond density between the primer terminal 3′-amino group and the substrate a-phosphate with con- comitant loss of electron density between the a- and b-phosphates with similar first-order rate constants (Fig. 1D). By inspecting the kinetics of density changes at nearby sites, we observed that the most prom- inent local density dynamics in the active site had simple linear relationships with nascent bond formation. Pairwise linear regression yields the relative slope (b) of density changes occurring at any two points across the time- resolved dataset (Fig. 2, A and B). Performing this regression at all points in the asymmetric unit versus the nascent bond furnished a field of b values or “beta map” in real space, which could be contoured at a desired threshold of |b| to give isosurfaces for visualization (Fig. 2C). To show that regions with high |b| are also highly correlated, we produced pairwise Pearson correlation maps using a similar procedure (Fig. 2D). Regions of local conformational dynamics were qualitatively in agreement between corre- lation maps with coefficient r contoured at |r| > 0.9 (Fig. 2D) versus regression maps contoured at |b| > 0.45 (Fig. 2C). When beta maps were Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E B A P O O O O O O O O O N N NH C D O P O O P O O P O Y654 D653 D830 NH2 Y710 dGTP b m u h T alm P Fingers Fig. 1. NP polymerase activ- ity proceeds in crystallo by a one-metal-ion mecha- nism. (A) Condensation of nNTPs generates NP-DNA. (B) Reaction scheme for BF-catalyzed primer extension of a 3′–amino-terminal primer on a DNA template, yielding pyrophosphate and an NP- linked +1 nucleotide extension. R = OH for in crystallo single- nucleotide addition of dGTP with co-crystallized divalent metal ion cofactor Ca2+, whereas R = NH2 for multiple turnover extension in solution to synthesize NP-DNA poly- nucleotide strands in the pres- ence of divalent (M2+) or trivalent (M3+) metal ion cofactors. (C) Ribbon cartoons of BF polymerase ground-state (GS) holocomplex x-ray crystal structure with 3′–amino-terminal DNA primer (green ribbon) and DNA template (cyan) with bound dGTP substrate and a single Ca2+ ion (green atom) in the canonical metal-B site (left and middle panels), showing one of two complexes co- crystallized in the asymmetric unit. All analyses performed here derive from the first complex (structure chains A to C), but consistent trends are seen in both. Dotted circle indicates the position of a missing A-site metal in analogous models of the canonical phosphodiester synthesis mechanism but not observed in closed conformation GS structures containing a 3′-NH2 (middle panel) or in product state (PS) structures of the in situ synthesized NP bond. (D) Representative time-dependent difference maps (Fo – Fc) showing density changes in the BF F710Y/D598A active site during NP bond formation in crystallo. GS crystals containing a 3′-amino- 2′,3′-dideoxycytidine (nC) terminated primer and DNA template were formed under acidic conditions (pH 5 to 6) in the presence of Ca2+ and dGTP, and the in situ reaction was initiated by transferring crystals to a solution at pH 8.8. The reaction in intact crystals was quenched at the indicated times by e g n a h C y t i s n e D d e z = 0.21±0.03 h-1 10 Time (h) Primer Ca2+ pH N-P P-O N-P K706 a m r o N t=0 D653 E831 nC P-O D830 1.0 0.0 0.6 0.8 0.4 0.2 20 i l 0 Template Primer O N H2 O O R BF O M2+/3+ O O P O P O OO O P O O R O O O P O O O N H O HO P OO O P O O pH A B GS PS t = 1 t= 2 t = 4 t=24h t=8 t = 6 flash freezing for subsequent data collection. Positive (aquamarine mesh) or negative (red mesh) isosurfaces are shown contoured at 2.5 s and superimposed on the GS model. Bottom left inset: normalized density change extracted from difference maps at the nascent N–P (aquamarine) and scissile O–P (red) bonds quantified from 19 crystals (dots). Observed first-order rate constant estimates, kNP and kPO, for the density changes ± SD are plotted (solid line ± shaded area). contoured at |b| > 0.9, we observed that the in- cluded voxels were tightly constrained to the site of the nascent NP bond and the scissile Pa–Oa,b bond (fig. S1A). No other linear density dynamics were larger in scalar magnitude than the chem- istry itself during bond formation in crystallo. Inspecting the isosurfaces contoured at |b| > 0.45, we observed four major density perturba- tions that were also highly correlated (|r| > 0.9) with the nascent bond (Fig. 2B, i to iv): (i) dis- placement of the terminal phosphodiester link- age between the –1 position and the primer terminus; (ii) a conformational switch of the substrate deoxyribose moiety from C2′-endo to C3′-endo, matching the sugar pucker of the primer terminus; (iii) a small deflection of the K706 side chain toward the substrate bridging a,b-oxygen, homologous to a putative general acid in the native mechanism (8); and finally (iv) disordering of the leaving group pyrophosphate moiety. As apparent from the high correlation coefficients, maps generated as 1 – p for two- tailed P values determined for the correlation field versus the nascent bond show that all of Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. One-metal-ion NP catalysis is limited by deficient substrate activa- tion. (A) Voxelwise linear regression analysis of density dynamics in crystallo. Difference map (Fo – Fc) densities derived from reacting crystals quenched at various times yield kinetic traces at a reference point (Ref.) or any query point of interest. A pairwise linear regression slope (b) of density changes at the reference versus query points can be estimated for any two points in real space. (B) Regression statistics of key density dynamics during NP synthesis in BF F710Y/D598A. Insets i to iv show difference map density derived from individual crystals at four major active-site blobs indicated in (C). Left panels: first-order kinetics of density differences from individual crystals (dots) over reaction time, with values for the estimated rate constant ± SD plotted (solid line ± shaded area). Right panels: pairwise linear regressions, coefficients of determination (R2), and correlation coefficients (r) of difference map densities at the indicated site versus the NP bond across all crystals. (C) Pairwise linear regression “beta” map calculated at all points versus the nascent NP bond (black wedge) with positive (aquamarine mesh) or negative (red mesh) b displayed as isosurfaces superimposed on the GS structure and contoured at |b| > 0.45. (D) Pearson correlation map of the active site contoured at |r| > 0.9 calculated by pairwise comparison of all voxels versus the peak of the nascent bond (black wedge), with positive (aquamarine mesh) or negative (red mesh) correlation displayed as isosurfaces superimposed on the GS structure. (E) Alternative view of the beta map contoured at |b| > 0.4, highlighting positive and negative beta peaks at sites (purple wedges) associated with a conformational change at the side chain of Lys706 [positive lobe quantified in (B), iii] in the direction of the substrate Oa,b bridging position and the scissile bond, as well as the conformation change in the dGTP substrate sugar [gray wedges, positive lobe quantified in (B), ii]. (F) Pairwise linear regression and correlation statistics versus the nascent bond calculated at a position equivalent to the absent A-site metal [gray “A” label in (C) and (D)]. (G) Cartoon overlay of GS (peach) and PS (blue) models with indicated conformational dynamics labeled as in (B) to (E). these active-site dynamics were highly signifi- cant with P < 10−6 (fig. S1B). Deflection of K706 toward the substrate a,b-bridging oxygen was well resolved on the beta map as positive and negative lobes surrounding terminal atoms of the side chain (Fig. 2E). This motion is note- worthy because the site is structurally homol- ogous to the C-site metal-ion density reported for X- and Y-family polymerases (5, 6, 7, 9). Crucially, no well-correlated density changes were observed at a position equivalent to the absent A-site metal in the native mechanism (Fig. 2F). Conformational changes in both the upstream and downstream nucleotides flanking the nas- cent bond were linearly correlated with covalent bonding (Fig. 2B, i and ii and C and D). The upstream conformational change at the primer- terminal linkage was also observable in time- resolved datasets of native phosphodiester synthesis in polymerase h (5). The downstream conformational change in the dGTP substrate sugar from C2′- to C3′-endo upon product formation (Fig. 2, B, inset ii, and E), however, was not observed in an equivalent time-resolved experiment with the wild-type enzyme, in which the conformation was stably C3′-endo (fig. S2). Although interesting, a ground-state confor- mational change in the deoxynucleotide sub- strate is unlikely to be relevant to 3′-amino Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E nucleotides, which already prefer the C3′- endo conformation due to a mixture of steric and anomeric effects (see the supplementary materials) (10). Isotopic and elemental substitution effects point toward rate-limiting chemistry with Ca2+ Reacting crystals were sampled on a time scale vastly slower than that of conformational dynamics such as furanose pseudorotation. The crystal data are therefore not directly informa- tive of the relative sequence and magnitude of these reaction barriers. To gain more insight into the rate-limiting step for NP synthesis, we turned to solution-phase kinetics. Two proton transfers have been detected in the rate-limiting step of native polymerase activity using careful measurements of the solvent deuterium kinetic isotope effect (SDKIE) (8, 11). In NP synthesis, it is minimally required that one proton arising from the 3′-amino nucleophile must ultimately be transferred out of the reaction center in the forward reaction, yielding the product phos- phoramidate. The conserved Asp830 side chain, proximal to the nucleophilic primer 3′-amine at ~2.5 Å in refined ground-state structures (Fig. 3B), is the most likely general base me- diating this proton transfer. A sterically con- servative point mutation at this position, D830N, entirely abolishes activity (1). We found that the effect of varying mole fraction, n, of D2O solvent on pre–steady-state NP reaction kinet- ics was negligible, yielding an SDKIE estimate of 1.16 for 3′-amino primer extension in the presence of Ca2+ and dCTP (Fig. 3A). In native phosphodiester polymerases, this value has generally been measured in the 2 to 5 range (8). The substantially lower value measured here suggests that proton transfer is not a criti- cal barrier in the NP reaction with Ca2+. If the chemical step of NP bond synthesis sets the overall reaction rate, then modulating the electrophilicity of the substrate should show an appreciable kinetic effect. We there- fore measured the nonbridging elemental thio substitution effect in solution using stereo- pure Sp or Rp diastereomers of dCTPaS. We found that the thio effect was kO/kS = 10.6 ± 0.5 (mean ± SD) when substituting the dCTP substrate with Sp-dCTPaS in pre–steady-state extension of a 3′-amino terminus at 45°C (Fig. 3, C and D). Elemental thio effects of similar magnitude have been observed for native poly- merase mismatch extension or when the di- valent cofactors are Mn2+ rather than Mg2+, both conditions for which the chemical step ap- pears to be rate limiting (11, 12). The thio effect was compounded by an additional fourfold for the Rp substrate versus the Sp substrate (Fig. 3D), consistent with the coordination geometry seen in crystal structures (Fig. 3B). The interpretation of thio effects has long been a matter of debate in phosphoryl transfer catalysis, given the potential for complicating A 1.2 1 0.8 0.6 0.4 0.2 C ° 5 5 @ O 2 H k / n k 0 0 C B nC Y710 d G T P pro-Sp 3.8 Å pro-Rp Å 5 . 2 D830 0.2 0.4 n 0.6 0.8 1 D2O E831 D653 3´NH2 F710Y Ca2+ +1,2,3... D dCTP Sp-dCTP S Rp-dCTP S 0 1 2 4 8 16 32 min C ° 5 4 @ ) 1 - n m i ( l o p k 0.08 0.06 0.04 0.02 0 1 0 . 6 ± 0 . 5 4.2 ± 0.7 dCTP Sp- S Rp- S Fig. 3. Probing for rate-limiting chemistry in NP-DNA synthesis in solution. (A) Pre–steady-state solvent deuterium kinetic isotope effect estimated from the slope of the line kn/kH2O = 1 + n(ϕ-1) for 3′-amino primer extension with 1 mM dCTP and 10 mM CaCl2 in varying mole fractions, n, of D2O. The quantity 1/ϕ is the SDKIE, estimated at 1.16 with 95% confidence interval (1.09 to 1.23) in the shaded region. (B) Atomic model distances in the GS reaction complex for the primer-terminal (nC) 3′-amine to the D830 side chain or to the substrate Pa. Gray mesh overlay indicates 2Fo – Fc density map contoured at 2.5 s. (C) Effect of a-phosphorothioate substitution on pre–steady-state Ca2+-activated NP synthesis at 45°C with 500 mM Sp-aS, Rp-aS, or unmodified dCTP substrates and BF F710Y. Inset: fluorescently labeled primer and template strand for the experiments in (A), (C), and (D). Representative 15% tris-borate EDTA (TBE)–urea PAGE separation of quenched reaction samples is shown. Only the first addition to this primer forms an NP or NPS linkage in the presence of 2′-deoxyribose substrates, yielding subsequent phosphodiester or phosphorothioate products, including mismatch extension products. (D) Elemental thio effects (labeled arrows, mean ± SD) estimated from the pre–steady-state rate constants (kpol, error bars indicate SD, n = 3) for reactions with substrates as in (C). B 2- O A N+1 O H O N O O O O E831 2+ M A O O D830 K706 OH O O NH3+ O O P OO M 2+ B O P O O O N H D653 Y654 P O O C 3- O K706 NH2 NH3+ O O P OO O O P O O M2+ P O O N+1 O H N H O N O O D830 O O O N H D653 Y654 N O O N+1 O H N H O D830 2- O K706 NH2 NH3+ O O P OO O O P O P O O M3+ O ? O O O N H D653 Y654 Di-divalent (3 -OH) Mono-divalent (3 -NH2) Mono-trivalent (3 -NH2) Fig. 4. Comparison of reaction models for phosphodiester and phosphoramidate synthesis in BF. (A) Model for the canonical two-metal ion “di-divalent” mechanism for native phosphodiester activity. (B) Model for the mono-divalent mechanism for Ca2+-catalyzed NP activity. (C) Proposed mono-trivalent mechanism for REE-catalyzed NP activity. Note that neither inner sphere waters nor exact proton-transfer pathways have been depicted. Two rate-limiting proton transfers have been detected for native activity (8, 11), but proton transfer is not rate limiting for mono-divalent NP activity with Ca2+. At least one proton must nevertheless be transferred to generate the phosphoramidate product from the attacking neutral amine. In BF, these transfers likely involve deprotonated D830 as a general base and protonated K706 as a general acid. Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 4 of 7 NH2 BF M3+, nNTPs 55 °C Sc3+ Lu3+ ~100 nt 1 - 2 hr Sc3+ +100 RES EARCH | R E S E A R C H A R T I C L E A C ° 5 5 @ ) 1 - n m i ( l o p k 8 7 6 5 4 3 2 1 0 101 100 10-1 10-2 10-3 C D 3´NH2 BF M3+, nCTP +1,2 B nCTP + Sc3+ 0 7 13 23 31 42 s Al3+ Ga3+ Er3+ La3+ Yb3+ Gd3+ Eu3+ +71 +70 Ca2+ Ca2+ Y3+ In3+ Lu3+ Sc3+ N.D. N.D. 3 2.5 2 1.5 1 C ° 5 4 @ ) 1 - n m i ( l o p k 0.5 0 3´NH2 BF, Sc3+ dCTP/dCTP S +1 0.58 ± 0.07 6 6 ± 7 dCTP Sp- S Rp- S dCTP Sp-dCTP S Rp-dCTP S 0 10 30 45 60 120 240 s 0 2 4 8 16 32 64 min 0 2 4 8 16 32 64 128 min Fig. 5. Trivalent rare earth metal ion cofactors confer rapid NP polymerase activity. (A) Pre–steady-state kpol estimates for extension of a 3′–amino- terminal DNA primer on a DNA template (inset cartoon) by BF F710Y/D598A at 55°C in the presence of various 1:1 ammonium citrate–buffered trivalent metal cations at 5 mM in 40 mM Tris-HCl, pH 8.8, 2 mM bME, in reactions initiated by addition of 250 mM nCTP. Estimates are displayed in linear (top) and log-scaled (bottom) axes compared with the level of Ca2+ activity (dotted line). Inset: representative 15% TBE-urea PAGE of quenched reaction samples from BF F710Y/D598A 3′–amino primer extension reactions in the presence of 5 mM Sc3+ showing +1 and +2 NP-DNA extension at the indicated times. ND, not detected. Error bars indicate SEM for n = 4 (Sc3+ and Lu3+) or n = 3 (In3+ and Y3+). (B) Mixed-sequence NP-DNA synthesis with excess BF F710Y/D598A on a 71-nt (left image) or 100-nt (right image) DNA-templating region at 55°C from a 5′–fluorescein-labeled 3′–amino-terminal DNA primer in the presence of 1:1 ammonium citrate–buffered 5 mM Sc3+ or Lu3+, as indicated, in 40 mM Tris-HCl, pH 8.8, 10 mM bME, and 25 mM spermine-HCl. Reactions were initiated by the addition of a 250 mM concentration of each nNTP, sampled and quenched at the indicated times, and samples were separated on 10% (left) or 8% (right) TBE-urea PAGE gels. (C) Elemental thio effects (labeled arrows, mean ± SD) for Sc3+-activated NP synthesis with 500 mM Sp-aS, Rp-aS, or unmodified dCTP substrates at 45°C estimated from kpol (error bars indicate SD, n = 3) using a fluorescently labeled primer and template strand (cartoon at top), yielding exclusively +1 product due to high NP(S) specificity. (D) Representative 15% TBE-urea PAGE separation of quenched samples for the reactions in (C). steric effects with phosphorothioate substrates (12, 13). Nevertheless, a plausible interpreta- tion of the observed thio effects is that the chemical step sets the overall reaction rate for NP synthesis. Electrophilic substrate activation as a probable deficiency in NP catalysis with Ca2+ If the chemical step is in fact rate limiting but proton transfer is not, then an augmented barrier to NP versus phosphodiester synthesis could, at least in part, originate from an altered metal-ion configuration in the presence of a 3′- amino primer. In the native reaction center, it is generally understood that the A-site metal ion activates the 3′-OH nucleophile by inner sphere coordination, yielding a metal alkoxide in the transition state, but numerous crystallo- graphic models of reaction intermediates also show that the octahedral ion bound at the A site forms an additional inner sphere contact with the substrate pro-Rp a-phosphate non- bridging oxygen (14). It has been indepen- dently argued that the role of the conserved polymerase dinuclear metal center (Fig. 4A) is to stabilize the transition state electrostatically on the basis of linear free energy relationships obtained from Brønsted plots of pre–steady- state kinetics in polymerase b, harnessing a series of Ob,g bridge substituent–modified sub- strate analogs with varying leaving group pKa (15, 16). The catalytic effect of the A-site metal in native activity likely encompasses several effects, but any catalytic effect conferred by the presence of an A-site metal ion on sub- strate activation, as opposed to nucleophile activation, is absent for NP synthesis. The lack of electron density for an A-site metal ion in the ground state and at later time points is also consistent with the expectation of a neu- tral nucleophile in the NP reaction and the generally weak affinity of aliphatic amines for Ca2+. Because neither the apparent bind- ing constant nor the active-site conforma- tion is significantly perturbed by the presence Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E of a 3′-amino nucleophile (1), a significant contributor to the kinetic defect for NP ver- sus phosphodiester synthesis may therefore be the relative loss of transition state stabi- lization associated with the missing A-site metal cofactor for the reaction with an amino nucleophile. If the native metal cofactors act, at least in part, on the native transition state by charge stabilization, then it is noteworthy that the net effect of the absence of one divalent ion (with disengagement of its conserved ligand, Glu831) would be to decrease the formal net charge of the transition state complex by 1 (Fig. 4, A and B). Outer sphere charge mutations may not be able to compensate for this inner sphere defect, par- ticularly if the role of the missing metal in native activity is partially to contribute to stabilizing the developing charge polarization across the scissile bond. This view is consistent with the observation that eliminating the neg- ative charge on the disengaged native ligand by the mutation E831Q fails to enhance NP synthesis kinetics (1). Trivalent metal ions confer rapid and specific NP synthesis activity On the basis of this activation argument and the observed thio effect, we hypothesized that substitution of a trivalent cation into the metal site B might compensate for the electrostatic component of the catalytic defect resulting from the absence of the A-site metal, following the hypothetical reaction structure depicted in Fig. 4C. Upon screening a series of redox- stable trivalent metal ions, we indeed found that an exotic series of trivalent metal ions could act as polymerase cofactors for rapid catalysis of NP bond formation (Fig. 5). The diamagnetic group three trivalent rare earth element (REE) cations scandium (Sc3+), yttrium (Y3+), and lutetium (Lu3+), as well as the post–transition metal ion indium (In3+), all significantly im- proved single-nucleotide 3′-amino primer exten- sion (Fig. 5A), as well as NP-DNA polymerization with all four 3′-amino 2′,3′-dideoxynucleoside 5′- triphosphates (nNTPs) on mixed-sequence templates (Fig. 5B and figs. S3 and S4). Sc3+ in particular accelerated pre–steady-state rate con- stants for nCTP addition by ~100-fold at 55°C to 7.1 ± 0.5 min−1 (mean ± SEM) versus 0.069 min−1 for Ca2+ (1), suggesting a stabilization effect of approximately –3 kcal/mol. Lu3+ yielded simi- lar levels of burst activity, kpol = 6.9 ± 0.7 min−1 (mean ± SEM), but had far weaker multiple turnover activity in long primer extensions rel- ative to Sc3+ (Fig. 5B and fig. S3). NP-active tri- valent ion cofactors have a wide range of pKa values for the aquo ion, are prone to hydrolysis under the reaction conditions, and have a ten- dency to precipitate nucleotides and other phos- phates. We found that reaction conditions were optimal when these ions were buffered at ~1:1 stoichiometry with citrate such that all reac- tion components remain homogeneously solu- ble (fig. S4). Although the reaction kinetics were quite sensitive to the metal:citrate stoichiome- try, they were not appreciably sensitive to the concentration of 1:1 Sc3+:citrate in the low- millimolar range (0.5 to 10 mM) in the presence of 1 mM total nNTPs (fig. S4). However, it is known that the metal-ligand stoichiometry of trivalent REE citrate complexes in solution is diverse (17). Trivalent metal ion cofactors confer exqui- sitely specific catalysis of NP versus phospho- diester chemistry. We found that the pre– steady-state rate constant for extension of a native DNA 3′-OH terminus with nCTP in the presence of Sc3+ was 5.5 ± 0.6 × 10−3 min−1, which is at least three orders of magnitude slower than 3′-NH2 extension (fig. S3B). NP-DNA synthesis activated by Sc3+ was also inhibited by added Mg2+ in the low millimolar range optimal for native activity (fig. S5A). The mutation D830N was completely inactivating, whereas the adjacent E831Q mutant exhibited wild-type activity in long extensions (fig. S5B), consistent with the active-site configuration proposed in Fig. 4C. Another notable feature of trivalent NP catalysis is an apparent in- verse elemental thio effect. We found that the Sc3+ thio effect was 0.58 ± 0.07 (mean ± SD) at 45°C, indicating a significant preference for phosphorothioate substrates to yield thio- phosphoramidate (NPS) linkages (Fig. 5, C and D). Inverse thio effects are more readily explained in the context of inverted stereo- specificity with thiophilic metal cofactor sub- stitution, but in this case the catalyzed reaction remained highly specific (66-fold) for the Sp over Rp substrates at 45°C. Synthesis of the potentially clinically relevant NPS linkages under these conditions was confirmed by high- resolution mass spectrometry (fig. S6). NP-DNA synthesis on long templates was aided by the addition of polyamines (fig. S4). With spermine and Sc3+, full-length products on mixed-sequence DNA templates up to +71 nucleotides (nt) or, with additional enzyme, up to +100 nt, could be synthesized in 1 to 2 hours at 55°C (Fig. 5B and fig. S4). Polyamine rescue is consistent with linkage-dependent confor- mational effects on duplex helicity for NP- DNA:DNA hybrids that arise during long-range primer extension synthesis, or it might reflect competition for inhibitory metal coordination sites. Activation by spermine was optimal at low micromolar concentrations, consistent with its reported effects on Klenow activity previously attributed to helicity effects (18). Under these conditions, NP-DNA synthesis remains highly determined by template se- quence, because omission of any single amino- nucleotide from the nNTP mixture markedly stalled extension activity at or immediately before the first template position complemen- tary to the missing substrate (fig. S7A and B). Extension products also show characteristic features expected from the incorporation of unnatural NP linkages, such as acid lability at elevated temperature (fig. S7A and B, HOAc lanes) and high resistance to 3′-5′ exonuclease I activity (fig. S7, C and D). Discussion Unlike native phosphodiester catalysis, only a single metal ion cofactor is detectable by crys- tallography during NP catalysis. Although tran- sition states and other short-lived intermediates were not directly observable in our studies, the totality of the evidence supports a single–metal ion mechanism with weakened substrate acti- vation in 3′-amino primer extension by BF. From this inference, we identified a series of trivalent metal ion polymerase cofactors with markedly enhanced reactivity and specificity for NP-DNA synthesis. Relative to the wild-type BF DNA polymerase, combining a single active- site point mutation with trivalent REE cofactor substitution accelerated NP synthesis by well over 1000-fold, yielding template-directed NP- DNA strands of useful lengths in benchtop enzyme reactions. In the presence of Sc3+ and spermine, overall NP polymerase reaction time to yield full-length products (~0.9 to 1.3 min/nt at 55°C) was similar to that re- ported for certain evolved xenonucleic acid (XNA) polymerases producing OP-linked strands incorporating synthetic sugars [0.75 to 1.5 min/nt at 50 to 65°C (19)]. REE-catalyzed NP polymerase reactions are therefore comparable to those of extensively mutated polymerases engineered for alternative phosphodiester synthesis activi- ties by directed evolution (20). This finding suggests that the evolutionary distance to ca- talysis of distinct chemistry may be negligible when alternative cofactors are present, given a conservatively modified substrate. By compari- son, recovering a conserved chemistry with highly divergent substrates may require a more extensive search of sequence space. The considerable catalytic advantage con- ferred by trivalent metal ion cofactors implies that a broader family of polymerase activities may be accessible by directed evolution of sub- strate specificity given the well-demonstrated ability of evolutionary methods to optimize native catalysis on XNA substrates (20, 21). NP-DNA falls within a class of phosphorami- date nucleic acids that have long been valued for their potential utility as nuclease-resistant antisense oligonucleotides (22). These poly- mers have also been extensively studied for their nonenzymatic self-assembly from chemi- cally activated phosphorimidazolide nucleo- tide analogs (23–26) on a path to developing synthetic protocells and model systems for key aspects of abiogenesis (27, 28). Despite lacking a 2′-OH group, short synthetic NP-DNA duplexes adopt a conformational geometry more similar to RNA than to DNA (10) such that Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E these polymers are viable templates for reverse transcriptase activity (29) but are poor sub- strates for several classes of nucleases (30, 31). Whether NP polymers are, like RNA, fully functional Darwinian polymers remains a mat- ter of considerable interest. The development of practical levels of NP polymerase activity is a crucial step on the path to demonstrating that this class of synthetic genetic materials is in fact functional and evolvable. RE FE RENCES AND N OT ES 1. V. S. Lelyveld, W. Zhang, J. W. Szostak, Proc. Natl. Acad. Sci. U.S.A. 117, 7276–7283 (2020). 2. T. S. Mansour, R. Storer, Curr. Pharm. Des. 3, 227–264 (1997). 3. A. K. Vashishtha, J. Wang, W. H. Konigsberg, J. Biol. Chem. 291, 20869–20875 (2016). 4. T. A. Steitz, Nature 391, 231–232 (1998). 5. T. Nakamura, Y. Zhao, Y. Yamagata, Y. J. Hua, W. Yang, Nature 487, 196–201 (2012). 6. B. D. Freudenthal, W. A. Beard, D. D. Shock, S. H. Wilson, Cell 154, 157–168 (2013). J. A. Jamsen et al., Nat. Commun. 8, 253 (2017). 7. 8. C. Castro et al., Nat. Struct. Mol. Biol. 16, 212–218 (2009). 9. Y. Gao, W. Yang, Science 352, 1334–1337 (2016). 10. V. Tereshko, S. Gryaznov, M. Egli, J. Am. Chem. Soc. 120, 269–283 (1998). 11. C. Castro et al., Proc. Natl. Acad. Sci. U.S.A. 104, 4267–4272 (2007). 12. C. M. Joyce, S. J. Benkovic, Biochemistry 43, 14317–14324 (2004). 13. D. Herschlag, J. A. Piccirilli, T. R. Cech, Biochemistry 30, 4844–4854 (1991). 14. W.-J. Wu, W. Yang, M.-D. Tsai, Nat. Rev. Chem. 1, 0068 (2017). 15. C. A. Sucato et al., Biochemistry 47, 870–879 (2008). 16. K. Oertell et al., Biochemistry 53, 1842–1848 (2014). 17. S. A. Wood, Eng. Geol. 34, 229–259 (1993). 18. A. Osland, K. Kleppe, Biochim. Biophys. Acta 520, 317–330 (1978). 19. A. I. Taylor, P. Holliger, Nat. Protoc. 10, 1625–1642 (2015). 20. N. Freund, M. J. L. J. Fürst, P. Holliger, Curr. Opin. Biotechnol. 74, 129–136 (2022). 21. A. A. Henry, F. E. Romesberg, Curr. Opin. Biotechnol. 16, 370–377 (2005). 22. S. Gryaznov et al., Nucleic Acids Res. 24, 1508–1514 (1996). 23. R. Lohrmann, L. E. Orgel, Nature 261, 342–344 (1976). 24. W. S. Zielinski, L. E. Orgel, Nucleic Acids Res. 13, 2469–2484 (1985). 25. S. Zhang, J. C. Blain, D. Zielinska, S. M. Gryaznov, J. W. Szostak, Proc. Natl. Acad. Sci. U.S.A. 110, 17732–17737 (2013). 26. W. Zhang, A. Pal, A. Ricardo, J. W. Szostak, J. Am. Chem. Soc. 141, 12159–12166 (2019). 27. S. S. Mansy et al., Nature 454, 122–125 (2008). 28. D. K. O’Flaherty, L. Zhou, J. W. Szostak, J. Am. Chem. Soc. 141, 10481–10488 (2019). 29. V. S. Lelyveld, D. K. O’Flaherty, L. Zhou, E. C. Izgu, J. W. Szostak, Nucleic Acids Res. 47, 8941–8949 (2019). 30. S. Gryaznov, J.-K. Chen, J. Am. Chem. Soc. 116, 3143–3144 (1994). 31. R. L. Hanna, S. M. Gryaznov, J. A. Doudna, Chem. Biol. 7, 845–854 (2000). ACKN OWLED GMEN TS We thank the staff at the Advanced Light Source (ALS) beamline 822 and the Advanced Photon Source (APS) beamline 23ID-B for technical assistance and F. Ng for invaluable research support. Funding: This work was supported by the Howard Hughes Medical Institute and Simons Foundation (grant 290363 to J.W.S.). The Berkeley Center for Structural Biology is supported in part by the Howard Hughes Medical Institute. The ALS is a Department of Energy (DOE) Office of Science User Facility supported by contract no. DE-AC02-05CH11231. The ALS-ENABLE beamlines are supported in part by the National Institutes of Health (NIH), National Institute of General Medical Sciences (grant P30 GM124169). GM/CA@APS has been funded by the National Cancer Institute (grant ACB-12002) and the National Institute of General Medical Sciences (grant AGM-12006, P30GM138396). This research used resources of the APS, a DOE Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under contract no. DE-AC02- 06CH11357. The Eiger 16M detector at GM/CA-XSD was funded by NIH grant S10 OD012289. Author contributions: V.S.L. and J.W.S. conceived the study and designed the research plan. Z.F. performed crystallizations, data collections, and refinements. V.S.L. analyzed the structural data, prepared materials, performed solution phase experiments, and analyzed the data. V.S.L. and J.W.S. interpreted the data and wrote the manuscript. Competing interests: A related patent application regarding enzymatic polymerization of NP-linked polynucleotides has been filed by V.S.L. and J.W.S. The remaining author declares no competing interests. Data and materials availability: Crystallographic data and maps have been deposited into the Protein Data Bank under PDB codes 8SCG, 8SCI, 8SCJ, 8SCK, 8SCL, 8SCM, 8SCN, 8SCO, 8SCP, 8SCQ, 8SCR, 8SCS, 8SCT, and 8SCU. All other data are presented in the main text or supplementary materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses- journal-article-reuse. This article is subject to HHMI’s Open Access to Publications policy. HHMI lab heads have previously granted a nonexclusive CC BY 4.0 license to the public and a sublicensable license to HHMI in their research articles. Pursuant to those licenses, the Author Accepted Manuscript (AAM) of this article can be made freely available under a CC BY 4.0 license immediately upon publication. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh5339 Materials and Methods Supplementary Text Figs. S1 to S7 Tables S1 to S3 References (32–39) MDAR Reproducibility Checklist Submitted 23 April 2023; accepted 20 September 2023 10.1126/science.adh5339 Lelyveld et al., Science 382, 423–429 (2023) 27 October 2023 7 of 7
10.1126_science.adh8160
RES EARCH BIOCHEMISTRY Structure of a ribonucleotide reductase R2 protein radical Hugo Lebrette1,2*†, Vivek Srinivas1†, Juliane John1, Oskar Aurelius1,3, Rohit Kumar1, Daniel Lundin1, Aaron S. Brewster4, Asmit Bhowmick4, Abhishek Sirohiwal1, In-Sik Kim4, Sheraz Gul4, Cindy Pham4, Kyle D. Sutherlin4, Philipp Simon4, Agata Butryn5,6‡, Pierre Aller5,6, Allen M. Orville5,6, Franklin D. Fuller7, Roberto Alonso-Mori7, Alexander Batyuk7, Nicholas K. Sauter4, Vittal K. Yachandra4, Junko Yano4, Ville R. I. Kaila1, Britt-Marie Sjöberg1, Jan Kern4*, Katarina Roos8§, Martin Högbom1* Aerobic ribonucleotide reductases (RNRs) initiate synthesis of DNA building blocks by generating a free radical within the R2 subunit; the radical is subsequently shuttled to the catalytic R1 subunit through proton-coupled electron transfer (PCET). We present a high-resolution room temperature structure of the class Ie R2 protein radical captured by x-ray free electron laser serial femtosecond crystallography. The structure reveals conformational reorganization to shield the radical and connect it to the translocation path, with structural changes propagating to the surface where the protein interacts with the catalytic R1 subunit. Restructuring of the hydrogen bond network, including a notably short O–O interaction of 2.41 angstroms, likely tunes and gates the radical during PCET. These structural results help explain radical handling and mobilization in RNR and have general implications for radical transfer in proteins. U ncontrolled or unmitigated free radicals can cause damage to cells but radicals are also essential to numerous meta- bolic pathways and enzyme-mediated chemistry (1, 2). Ribonucleotide reduc- tase (RNR) is an archetypal radical enzyme, and the tyrosyl radical (Y(cid:129)) in the R2 subunit from Escherichia coli was the first stable pro- tein radical to be observed, 50 years ago (3). RNR provides the only pathway for de novo synthesis of deoxyribonucleotides and repre- sents a drug target for both cancer and infec- tious diseases (4, 5). Aerobic RNR (class I) depends on the ferritin-like R2 subunit to gen- erate a catalytic radical in an oxygen-dependent reaction; the radical must be transferred back and forth with the catalytic R1 subunit, which performs the ribonucleotide reduction (6). Radi- cal translocation between the subunits proceeds through reversible long-range proton-coupled electron transfer (PCET) in a transient R1-R2 complex (7). Radical transfer initiation in- volves redox-induced structural changes in R2 (8, 9), conformational gating (10), short-range 1Department of Biochemistry and Biophysics, Stockholm University, Arrhenius Laboratories for Natural Sciences, Stockholm, Sweden. 2Laboratoire de Microbiologie et Génétique Moléculaires, Centre de Biologie Intégrative, CNRS, Université Toulouse III, Toulouse, France. 3MAX IV Laboratory, Lund University, Lund, Sweden. 4Molecular Biophysics and Integrated Bioimaging Division, Lawrence Berkeley National Laboratory, Berkeley, CA, USA. 5Diamond Light Source Ltd, Harwell Science and Innovation Campus, Didcot, UK. 6Research Complex at Harwell, Harwell Science and Innovation Campus, Didcot, UK. 7LCLS, SLAC National Accelerator Laboratory, Menlo Park, CA, USA 8Department of Cell and Molecular Biology, Uppsala University, Uppsala, Sweden. *Corresponding author. Email: hugo.lebrette@univ-tlse3.fr (H.L.); jfkern@lbl.gov (J.K.); hogbom@dbb.su.se (M.H.) †These authors contributed equally to this work. ‡Present address: Macromolecular Machines Laboratory, The Francis Crick Institute, London, UK. §Present address: Sprint Bioscience, Huddinge, Sweden. proton transfer (11) coupled to long-distance electron transfer (12), and regulation by R1 (7). Recently, the structure of an R1-R2 holo- complex was determined by cryo–electron mi- croscopy (cryo-EM) (7) providing a picture of the long-range radical transfer pathway. How- ever, atomic resolution snapshots of the radi- cal state and the conformational gating taking place at the R2 active site remain unresolved. Most R2 proteins harbor a conserved tyro- sine residue oxidized to Y(cid:129) by an oxygen- activated metal center in the catalytically active state. In a recently discovered active metal-free R2 subclass, denoted R2e, this tyrosine residue is post translationally meta-hydroxylated to a 3,4-dihydroxyphenylalanine (DOPA) which serves as the radical-harboring residue (13, 14). Oxygen activation of R2e and metal-containing R2 pro- teins display analogous pathways consisting of a nonactivated state, a catalytically active radical state, and a radical-lost ground state (Fig. 1A). From an experimental point of view, R2e from Mesoplasma florum (MfR2) represents an attractive model to study active radical states in RNRs. Its radical state is metal-independent which simplifies sample preparation and ex- cludes partial occupancy or mismetalation of the metal site often encountered with metallo- enzymes in vitro (15). In addition, in absence of protein R1 its radical state is stable with no decay observed after more than six hours at 25°C (13). However, as with any radical state, it is expected to be highly sensitive to photoreduction as x-ray radiation damage generates free radicals that spread through protein crystals (16). This obstacle renders the structural characteriza- tion of a protein radical state using standard x-ray crystallography methods unfeasible (17, 18). Though many crystal structures of iron- and manganese-containing R2 proteins from dif- ferent organisms have been obtained, they describe either the nonactivated state, the radical-lost ground state (often referred to as the “met” state) or partially reduced states. Crys- tal structures of R2e have been solved from two different organisms (13, 14), likewise showing signs of x-ray–induced photoreduction. Discrepancies between R2 crystal structures and spectroscopic data of radical states have been observed, leading to contradictory theo- retical models regarding the conformation of Y(cid:129) and its environment (8, 19–21). In the present study, by rapid protein production, microcrys- tallization and x-ray free-electron laser (XFEL) serial femtosecond crystallography, using the diffraction-before-destruction principle (22), we have determined the atomic structure of MfR2 in the active radical state at room tem- perature. Compared with the structure of the radical-lost ground state, also presented here, the radical state structure reveals a notably short hydrogen bond and a critical rearrange- ment of conserved residues upon acquisition of the radical. Based on these two distinct states, we propose a mechanism for structural recog- nition of the radical state and a model for redox-coupled conformational gating as a pro- logue to the radical transfer. This mechanism defines central aspects of the PCET process and may be conserved in aerobic RNRs. Structure of the radical-lost ground state of MfR2 In the catalytically active radical state, the radical-harboring meta-hydroxylated tyrosyl (from here denoted DOPAY126(cid:129)) in MfR2 exhib- its a characteristic absorbance peak at 383 nm and colors the protein blue (13). The catalytic radical can be chemically quenched by hydro- xyurea, a known RNR radical scavenger used for decades as an antitumor drug (23) pro- posed to inactivate R2 through PCET (24). Im- portantly, hydroxyurea causes a reversible inactivation of MfR2 as the enzyme can recover activity upon reoxidation by NrdI (13). Thus, the protein is not permanently inactivated or damaged but resides in a radical-lost ground state (Fig. 1A). To ensure an accurate depiction of the radical-lost ground state of MfR2, we solved the crystal structure at 1.35 Å resolution of the pro- tein chemically quenched by incubation with hydroxyurea before crystallization. This treat- ment abolished the 383-nm absorbance peak and rendered the protein colorless (13). The electron density map clearly shows the post translational modification in the meta posi- tion of DOPAY126 (Fig. 1B). The two oxygen- containing functional groups of DOPAY126 in the para and meta positions (denoted para-O and meta-O, respectively) form hydrogen bonds (H-bond) with D88. Notably, the meta-O is in- volved in an unusually short interaction (Fig. 1B). Using END/RAPID error analysis (25), the O–O distance between DOPAY126 and D88 was Lebrette et al., Science 382, 109–113 (2023) 6 October 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Radical state and radical-lost ground state of MfR2. (A) Outline of the proposed activation pathway of metal-free R2e, including the nonactivated state, catalytically active radical state, and radical-lost ground state (also known as “met” in canonical R2). The two states determined in this work are indicated in green. The structure of R2e in the nonactivated state has been determined previously (13, 14). For clarity, chemical reactions are not strictly balanced. (B) Structure of MfR2 in the radical-lost ground state obtained after chemical quenching by hydroxyurea (monomer A is shown). (C) Structure of MfR2 in the radical state obtained from XFEL serial femtosecond crystallography showing a reorganization of the site compared with the ground state including a coupled movement of the DOPAY126-D88 dyad, the presence of a new water w1, and the inward conformation of K213. The short H-bond is highlighted in orange. (D) Superimposition of the ground and radical states. The 2-Å displacement of the DOPA para-O is marked in purple. Nitrogen and oxygen atoms are shown in blue and red, respectively. Carbons are shown in gray and cyan for the radical- lost ground state and radical state, respectively. Distance between atoms involved in H-bond interactions are in Å. Simulated annealing composite Omit 2Fo−Fc electron density maps are shown in green and contoured at 2 s. The structural changes are further illustrated in movie S1. (E and F) Using quantum mechanical calculations on the XFEL structure, the short H-bond between DOPAY126 and D88 can be reproduced by a DOPA radical state with the radical mainly located on the para-O and the proton located on meta-O (E) (neutral DOPA radical state) or D88 (F) (negatively charged DOPA radical state). For clarity, only a subset of residues included in the calculations is presented on the figure (see fig. S3 for full details). calculated to 2.43 ± 0.04 Å, a length which could correspond to a low-barrier or single- well H-bond (26, 27). In addition, DOPAY126 does not interact with any water, nor with K213 whose e-ammonium group is facing away from DOPAY126 (Fig. 1B). Compared with previous R2e structures, differences and al- ternate conformations are observed through- out the structures, particularly variations in the conformation of DOPAY126, D88, and K213 (fig. S1A). Structure of the radical state of MfR2 by serial femtosecond crystallography To circumvent photoreduction artifacts, we used XFEL serial femtosecond crystallography to determine the structure of the catalytically active radical state of MfR2 at atomic reso- lution. The active radical-harboring protein was batch crystallized to produce a suspension of blue microcrystals used to collect room tem- perature serial femtosecond diffraction data at an XFEL source. The resulting dataset produces a model of the protein at 1.5-Å resolution, and the electron density map allows unambiguous interpretation (Fig. 1C). A short H-bond be- tween DOPAY126 meta-O and D88 is present with a O–O distance calculated to 2.41 ± 0.05 Å using END/RAPID error analysis (25) (see Methods for details). A short H-bond is ob- served in both states of the protein and may play a special structural role as suggested in other cases (28), contributing to maintaining the integrity of the enzyme active site (29). The short H-bond may stabilize the interaction between DOPAY126 and D88 to ensure that no hydrogen is available to mediate a putative pro- ton transfer from D88 to the DOPAY126 para-O, which would annul productive PCET. We note that this H-bonding structure results in a sit- uation analogous to canonical R2 proteins in which a deprotonated aspartate is involved in metal coordination and not in proton transfer, thus forcing the latter to occur with a different nearby proton donor, suggested to be a metal- bound water molecule (11). Furthermore, it is tempting to speculate that this short H-bond is involved in redox tuning. By preventing the radical delocalization between the meta-O and para-O, the DOPA radical becomes electroni- cally more similar to a tyrosyl radical, rather than a DOPA-semiquinone radical, in agreement with previous characterization of MfR2 by electron paramagnetic resonance (EPR) (13). A notable conformational shift takes place upon radical acquisition between DOPAY126 and D88: the dyad undergoes a coupled coplanar (but not coaxial) rotation of ~22°, with an ad- ditional rotation of the aromatic ring of ~12° along the Cb-C1 axis. It results in a 2-Å dis- placement of the para-O carrying the main radical spin density away from D88 and a re- organization of the interaction pattern around DOPAY126 (Fig. 1D). The aspartate residue cor- responding to D88 in MfR2 is strictly con- served as the N-terminal metal-coordinating residue in Y(cid:129)-harboring canonical R2 proteins and exhibits redox-induced conformational changes in metal-containing R2 proteins (8, 17, 30, 31). Furthermore, coupled movements of the conserved radical-harboring Y have been observed to be redox-induced in R2b from Bacillus anthracis (17). For R2a from E. coli (EcR2a), the conserved aspartate is proposed to form a H-bond with the reduced Y in the ground state. By contrast, single-crystal EPR experiments suggest that the Y(cid:129) rotates away in the radical state, leading to a ~1-Å displace- ment of the radical-harboring oxygen and breaking the connection with the aspartate (8). A displacement of the Y(cid:129) could also be Lebrette et al., Science 382, 109–113 (2023) 6 October 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E hypothesized from discrepancies observed between crystal structures and spectroscopic data for R2 proteins from Bacillus anthracis (32), Salmonella typhimurium (33), Coryne- bacterium ammoniagenes (30) and mouse (19). This type of rearrangement upon acquisition of the radical is principally similar to what we observe in MfR2 structures and is less pronounced than movements proposed in other studies, which involve either translation of the main chain or larger Y(cid:129) displacement by several Å (20, 21, 24). In addition, a clearly defined water molecule (w1) mediates a new H-bond between DOPAY126 and the e-ammonium group of K213, which displays a different orientation facing toward DOPAY126 (Fig. 1C). K213 adopts a single well- defined conformation different from previously published structures of active R2e determined using synchrotron radiation (13) (fig. S1B). The presence of a water molecule at a position sim- ilar to w1 has been observed previously in other RNR systems (34, 35), and seems to be de- pendent on the redox state (17) (fig. S2). More- over, superimposition with structures of R2 proteins shows that the new location of the K213 e-ammonium group corresponds to the position of another water molecule that is metal- coordinated in canonical R2 proteins (fig. S2). This water is proposed to transfer a proton to Y(cid:129) in the conformational gate initiating the PCET (11, 36, 37). Residue K213 was recently suggested by density functional theory to be a proton donor for radical transfer in R2e (38). There- fore, it may represent the water-equivalent proton donor in the case of R2e, and thus its conformational change could effectuate a com- parable conformational gate. Comparing the structures of the defined rad- ical and radical-lost states determined here to previously solved structures of R2e proteins shows that no previous structure fully repre- sents either the radical state or the radical- lost state (fig. S1). Although the proteins in prior work may have originally crystallized in the “active form,” they appear to have suffered different degrees of x-ray–induced photo- reduction during synchrotron data collection. The XFEL structure can be reproduced in silico by a radical state To evaluate whether the crystal structure ob- tained by XFEL femtosecond crystallography theoretically corresponds to a radical state, calculations were performed on the crystal struc- ture active site. Based on quantum chemical geometry optimizations, the short interaction between DOPAY126 and D88 could be reproduced with the main radical character on the para-O. The proton could reside on either DOPAY126 or D88 as both states produced a short O–O dis- tance (Fig. 1, E and F, and fig. S3, A and B). The calculated energy difference of 3 kcal/mol be- tween the two states suggests that both states are accessible with slight favoring of the proton residing closer to DOPAY126(cid:129). In addition, var- ious alternative DOPAY126 states were modeled by quantum mechanical calculations, none of them agreeing well with the experimental ob- servations. In particular, a longer hydrogen bond with D88 is observed when DOPAY126 is modeled as a neutral DOPA, a DOPA quinone, or with the radical located on the meta-O (fig. S3, C to E). Furthermore, the calculated spin population of DOPAY126(cid:129) in the protein active site models revealed an asymmetric distribution closer to a meta-substituted Y(cid:129) than the fully delocalized character of an ortho-semiquinone, based on comparisons with calculated distributions in smaller models. This is fully consistent with spectroscopic data of the radical (13, 14) and indicates a possible role for the short DOPAY126- D88 hydrogen bond to destabilize the other- wise potentially too stable semiquinone radical in the protein. Molecular dynamics simulation starting from the XFEL structure of the radical state but with induced loss of the radical in silico, showed DOPAY126 movement with the dynamics domi- nated by the position of the radical-lost ground state, forming two H-bonds to D88 (fig. S3, F to H), consistent with the crystal structure. Altogether, our calculations support that the structure obtained by XFEL corresponds to the catalytically active radical state of the MfR2 protein and are in agreement with previous EPR and UV–vis spectroscopic results (13), showing that the spin density distributes sim- ilar as in a meta-substituted tyrosyl radical ra- ther than as in an ortho-semiquinone. Specific protein rearrangements upon radical acquisition The radical acquisition in MfR2 leads to two major protein rearrangements that are of par- ticular interest as they can be directly impli- cated in radical generation, stabilization, and transfer. The first major protein rearrangement takes place within the activating-oxidant path. The channel connecting the NrdI flavin cofactor to the R2b metal site (39, 40), proposed as the (cid:129)− route for activation, seems to be con- O2 served in R2e (14). In the radical-lost ground state of MfR2, in place of the metal center present in canonical R2, a continuous chain of well-defined H-bonded water molecules cre- ates the link between the putative oxidant route and the DOPAY126-D88 dyad. By contrast, in the radical state of MfR2, this water network is disrupted by Q91 which undergoes a large sidechain flip toward D88 (Fig. 2 and movie S1). In the radical-lost ground state, Q91 is involved in a H-bond network conserved in the R2b subclass (Q70 in R2b from E. coli), which lines the channel for oxidant transport to the MnII 2 active site (40). Our data suggest that Q91 could play a key role as it obstructs the putative oxidant channel in the radical state, preventing radical quenching by further oxi- dant access to the active site. The second major protein rearrangement upon radical acquisition concerns the PCET route. In the immediate vicinity of the side chain of DOPAY126, residues L183 and F187 undergo large conformational changes in the radical Fig. 2. Major conformational changes between the radical-lost ground state and the radical state. (A) Structure of MfR2 in the radical-lost ground state obtained after chemical-quenching by hydroxyurea. (B) Structure of MfR2 in the radical state obtained from XFEL serial femtosecond data showing reorganization of the site compared with the radical-lost ground state, including conformational changes of Q91, L183, and F187. In the radical state, Q91 displaces two water molecules, breaking the water H-bond network toward DOPAY126. Structural movements of L183 and F187 leave space for 3 water molecules interacting with D88, DOPAY126, and K213. Simulated annealing composite Omit 2Fo−Fc electron density maps are shown in green and contoured at 1.5 s. (C) Superimposition of the structures of MfR2 in the radical state (cyan) and radical-lost ground state (gray). Lebrette et al., Science 382, 109–113 (2023) 6 October 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E state of MfR2, leaving space for the position- ing of three water molecules (including w1) and creating a water-mediated H-bond net- work between DOPAY126, D88, K213, and Q91 (Fig. 2 and movie S1). Residues L183 and F187 belong to helix aE which forms a distorted p-helix conserved across many ferritin super- family members. The p-helix conformation of helix aE is believed to play a functional role (41, 42), and is known to undergo redox-induced structural rearrangements (17, 30, 34). Resi- dues L183 and F187 are conserved in all R2 proteins, the position of L183 being either F or L (in 91 and 9% of sequences, respectively) and F187 being conserved in 99% of the se- quences. In none of the R2 crystal structures solved to date do these residues exhibit con- formations resembling those in the MfR2 rad- ical state (fig. S4). In the first solved structure of R2, it was noted that the radical-harboring tyrosyl oxygen was surrounded by a conserved hdrophobic pocket formed by residues F208, F212, and I234 in EcR2a (equivalents to L183, F187, and I209 in MfR2, respectively) (43). The major function of these residues was proposed through mutational studies to contribute to the tyrosyl radical stability by insulating the radical-harboring tyrosyl oxygen (44). Our ob- servation of hitherto unseen movements of these radical-shielding residues further impli- cates them in radical control and suggests their involvement in gating of the ribonucleotide reductase PCET mechanism. The specific local rearrangements at the radical site also trans- late to movements of the protein backbone and global structural changes of the protein scaffold protruding to the R1 interaction surface and the radical transfer path (movie S2). It has previ- ously been shown that the active R1-R2 complex exhibits tighter binding after radical initiation (45–47). We propose that these global structural changes observed in R2 provide a mechanism by which the R1-R2 binding properties can be modulated during the catalytic cycle. Model of conformational gating for radical transfer initiation In MfR2, the catalytically active radical state and the radical-lost ground state are interconvertible by quenching the radical and through NrdI- mediated reoxidation of DOPAY126 (Fig. 1A). Based on our structural data, we propose a model of the conformational gating orches- trated by R2 after radical acquisition, which is a prelude to the radical translocation to the R1 subunit (Fig. 3). This model proceeds in three steps. First, the oxidation of DOPAY126 leads to a repulsion between its para-O and D88 due to the removal of the H-bonding hydro- gen atom, resulting in the 2-Å displacement of the DOPAY126 para-O (Fig. 3A). Secondly, this triggers a cascade of structural changes to shield the radical and prepare its transfer: Q91 blocks the access to further oxidant, L183 and F187 reshape the insulating pocket around the radical, and water w1 connects the DOPAY126 radical-carrying oxygen with the PCET route (Fig. 3B). This also leads to global structural rearrangements of protein R2, including the R2-R1 interaction surface and binding of protein R1 (Fig. 3C and movie S2). Thirdly, the formation of the R1-R2 complex results in the ordering of the full R2 C-terminal tail at the R1-R2 interface [as demonstrated in (7)], completing the electron transfer path and in- ducing the injection of an electron to reduce DOPAY126(cid:129), e.g., through the conserved W52 and/or Y325 (corresponding to W48 and Y356 in EcR2a), as previously suggested (12) (Fig. 3D). Coupled to this event, a proton transfer from K213 to DOPAY126(cid:129) occurs by means of the water w1 [as proposed in (38)] and initiates the long- range radical translocation. Figure 3 summarizes the proposed steps of conformational gating in R2e. The interaction with R1 is modeled based on a superposition of the cryo-EM structure of the R1-R2 holocomplex from E. coli (7). This conformational gating model for ini- tiating the radical translocation could be com- mon to class I RNR systems. In PCET, proton Fig. 3. Model of conformational gating and radical transfer during PCET in R2e. (A) Oxidation of the radical-lost ground state DOPAY126 (gray) displaces DOPAY126-D88 and introduces a water molecule (w1) to facilitate the proton transfer from the redirected amino group of K213 (cyan). (B) Concurrently, the cavity surrounding the generated DOPAY126 radical is reshaped, shown in surface representation (radical-lost state and radical state in gray and cyan, respectively). Primarily, Q91 flips away from Y163 to block access of any further oxidant from NrdI through the R2-NrdI channel and L183 and F187 flip to reshape the cavity and facilitate electron transfer from W52 to DOPAY126. (C) Structure superposition of MfR2 in the radical (cyan) and radical-lost (gray) states at the R1 binding interface. Reshaping of the pockets surrounding the DOPAY126 leads to conformational changes protruding to the R2-R1 interaction surface and binding to protein R1 completes the PCET pathway. (D) PCET is initiated by a proton transfer (PT) from the amino group of K213 to DOPAY126 mediated by w1, and simultaneously an electron transfer from W52 to DOPAY126, thus trans- locating the radical toward the active site C394 of R1. R1 (represented in red) is modeled in complex with the radical-state MfR2 using the cryo-EM structure of the R1-R2 holocomplex from E. coli (PDB ID: 6w4x) (7). Lebrette et al., Science 382, 109–113 (2023) 6 October 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E transfer occurs through H-bond networks and requires the proton donor and acceptor to be within a standard ~2.8-Å H-bond distance (48). As a displacement of the radical-harboring Y upon radical acquisition is also observed by spectroscopy in other R2 proteins (8, 19, 30, 32, 33), an intermediary may be required between the metal-bound water and Y(cid:129). In MfR2 the ad- ditional water w1 gained upon radical acquisi- tion represents the missing piece of the puzzle, connecting the radical to K213 (Fig. 3D). We note that an analogous binding site for water has been observed in other R2 proteins (fig. S2), and that the e-ammonium group of K213 is located at the position of the metal-bound water in metal-containing R2s (fig. S2). There- fore, we speculate that, similarly as in R2e, a water in this position links the radical to the metal-bound water proposed to be the proton donor in canonical R2 proteins and gates the first proton transfer to initiate radical translo- cation to R1. RE FE RENCES AND N OT ES 1. M. Högbom, B.-M. Sjöberg, G. Berggren, in eLS (Wiley, 2020), vol. 1, pp. 375–393. J. Stubbe, D. G. Nocera, J. Am. Chem. Soc. 143, 13463–13472 (2021). 2. 3. A. Ehrenberg, P. Reichard, J. Biol. Chem. 247, 3485–3488 (1972). 4. L. Miret-Casals et al., ACS Omega 3, 17057–17069 (2018). 5. Y. Xie et al., J. Biomed. Sci. 29, 32 (2022). 6. R. Banerjee, V. Srinivas, H. Lebrette, in Macromolecular Protein Complexes IV, J. R. Harris, J. Marles-Wright, Eds. (Springer, 2022), vol. 99 of Subcellular Biochemistry, pp. 109–153. 7. G. Kang, A. T. Taguchi, J. Stubbe, C. L. Drennan, Science 368, 424–427 (2020). 8. M. Högbom et al., Proc. Natl. Acad. Sci. U.S.A. 100, 3209–3214 (2003). 9. A. R. Offenbacher, I. R. Vassiliev, M. R. Seyedsayamdost, J. Stubbe, B. A. Barry, J. Am. Chem. Soc. 131, 7496–7497 (2009). 10. K. Yokoyama, U. Uhlin, J. Stubbe, J. Am. Chem. Soc. 132, 15368–15379 (2010). 11. B. Wörsdörfer et al., J. Am. Chem. Soc. 135, 8585–8593 (2013). 12. S. Y. Reece, M. R. Seyedsayamdost, Essays Biochem. 61, 281–292 (2017). 13. V. Srinivas et al., Nature 563, 416–420 (2018). 14. E. J. Blaesi et al., Proc. Natl. Acad. Sci. U.S.A. 115, 10022–10027 (2018). 15. J. A. Cotruvo Jr., J. Stubbe, Metallomics 4, 1020–1036 (2012). 16. E. F. Garman, Acta Crystallogr. D 66, 339–351 (2010). 17. K. Grāve et al., J. Biol. Inorg. Chem. 24, 849–861 (2019). 18. K. G. V. Sigfridsson et al., J. Biol. Chem. 288, 9648–9661 (2013). 19. K. R. Strand et al., J. Biol. Chem. 279, 46794–46801 (2004). 20. B. A. Barry et al., J. Phys. Chem. Lett. 3, 543–554 (2012). 21. A. R. Offenbacher, L. A. Burns, C. D. Sherrill, B. A. Barry, J. Phys. Chem. B 117, 8457–8468 (2013). 22. H. N. Chapman et al., Nature 470, 73–77 (2011). 23. N. Saban, M. Bujak, Cancer Chemother. Pharmacol. 64, 213–221 (2009). 24. A. R. Offenbacher, B. A. Barry, J. Phys. Chem. B 124, 345–354 (2020). 25. P. T. Lang, J. M. Holton, J. S. Fraser, T. Alber, Proc. Natl. Acad. Sci. U.S.A. 111, 237–242 (2014). 26. C. L. Perrin, J. B. Nielson, Annu. Rev. Phys. Chem. 48, 511–544 (1997). 27. C. L. Perrin, Acc. Chem. Res. 43, 1550–1557 (2010). 28. D. A. Nichols et al., J. Am. Chem. Soc. 137, 8086–8095 (2015). 29. M. T. Kemp, E. M. Lewandowski, Y. Chen, Biochim. Biophys. Acta. Proteins Proteomics 1869, 140557 (2021). 30. N. Cox et al., J. Am. Chem. Soc. 132, 11197–11213 (2010). 31. D. T. Logan et al., Structure 4, 1053–1064 (1996). 32. E. Torrents, M. Sahlin, D. Biglino, A. Gräslund, B.-M. Sjöberg, Proc. Natl. Acad. Sci. U.S.A. 102, 17946–17951 (2005). 33. M. Galander, M. Uppsten, U. Uhlin, F. Lendzian, J. Biol. Chem. 281, 31743–31752 (2006). 34. M. Eriksson, A. Jordan, H. Eklund, Biochemistry 37, 13359–13369 (1998). 35. M. Högbom, Y. Huque, B.-M. Sjöberg, P. Nordlund, Biochemistry 41, 1381–1389 (2002). 36. P. E. M. Siegbahn, L. Eriksson, F. Himo, M. Pavlov, J. Phys. Chem. B 102, 10622–10629 (1998). 37. J. Stubbe, D. G. Nocera, C. S. Yee, M. C. Y. Chang, Chem. Rev. 103, 2167–2201 (2003). 38. J. Zou, Y. Chen, W. Feng, Comput. Struct. Biotechnol. J. 20, 1111–1131 (2022). 39. J. John et al., eLife 11, e79226 (2022). 40. A. K. Boal, J. A. Cotruvo Jr., J. Stubbe, A. C. Rosenzweig, Science 329, 1526–1530 (2010). 41. D. Lundin, A. M. Poole, B.-M. Sjöberg, M. Högbom, J. Biol. Chem. 287, 20565–20575 (2012). 42. R. B. Cooley, D. J. Arp, P. A. Karplus, J. Mol. Biol. 404, 232–246 (2010). 43. P. Nordlund, B. M. Sjöberg, H. Eklund, Nature 345, 593–598 (1990). 44. M. Ormö et al., J. Biol. Chem. 270, 6570–6576 (1995). 45. B. L. Greene, D. G. Nocera, J. Stubbe, J. Biol. Chem. 293, 10413–10414 (2018). 46. E. C. Minnihan et al., Proc. Natl. Acad. Sci. U.S.A. 110, 3835–3840 (2013). 47. B. L. Greene et al., Annu. Rev. Biochem. 89, 45–75 (2020). 48. A. Migliore, N. F. Polizzi, M. J. Therien, D. N. Beratan, Chem. Rev. 114, 3381–3465 (2014). 49. H. Lebrette, K. Roos, V. Kaila, M. Högbom, QM/MM and MD result files for the MfR2e system, SciLifeLab Data Repository (2023). ACKN OWLED GMEN TS We thank R. Bolotovsky and I. D. Young for help with data processing and the staff at LCLS. We acknowledge the Paul Scherrer Institute, Villigen, Switzerland for provision of synchrotron radiation beamtime (proposals 20161653 and 20182304), and the staff of beamline X06SA (PXI) at the Swiss Light Source for assistance. We acknowledge SOLEIL for provision of synchrotron radiation facilities (proposal 20180270) and we thank the staff of beamlines PROXIMA 1 and PROXIMA 2A for assistance. Research was carried out at the Linac Coherent Light Source (LCLS), SLAC National Accelerator Laboratory (proposal LU50), supported by the DOE Office of Science, OBES under contract DE-AC02-76SF00515. XFEL data processing was performed in part at the National Energy Research Scientific Computing Center, supported by the DOE Office of Science, contract DEAC02- 05CH11231. The Rayonix detector used at LCLS was supported by the NIH grant S10 OD023453. Experiments at the LCLS were supported by the NIH grant P41GM139687. D.L. acknowledges the Centre for Ecology and Evolution in Microbial model Systems – EEMiS, Linnaeus University. A.S. and V.R.I.K. acknowledge the National Academic Infrastructure for Supercomputing in Sweden (NAISS 2023/1-31) for computational resources. Funding: The authors acknowledge financial support of this work by the Swedish Research Council (2021-03992 to M.H. and 2019-01400 to B.-M.S.), by the European Research Council (HIGH-GEAR 724394 to M.H.), by the Knut and Alice Wallenberg Foundation (2017.0275 and 2019.0436 to M.H., 2019.0251 to V.R.I.K.), by the Swedish Cancer Foundation (20 1210 PjF to B.-M.S.), by the Swedish Foundation for Strategic Research (ICA16-0040 to K.R.), by the Swedish National Infrastructure for Computing (SNIC 2021/5-137 to K.R.), by the Director, Office of Science, Office of Basic Energy Sciences (OBES), Division of Chemical Sciences, Geosciences, and Biosciences (CSGB), Department of Energy (DOE) (to J.Y., V.K.Y., and J.K.), by the National Institutes of Health (NIH) grants GM133081 (to K.D.S.), GM055302 and 1R35GM149528-01 (to V.K.Y.), GM110501 (to J.Y.) GM126289 (to J.K.), GM117126 (to N.K.S.), by Diamond Light Source, the UK Science and Technology Facilities Council (STFC), and Biotechnology and Biological Sciences Research Council (to A.M.O., P.A., and A.Bu.), by a Wellcome Investigator Award (210734/Z/18/Z to A.M.O.), by a Royal Society Wolfson Fellowship (RSWF\R2\182017 to A.M.O.), and by an EMBO Long-Term Fellowship (ALTF 952-2022 to A.S.). Author contributions: H.L. performed crystallography, processed and analyzed synchrotron and XFEL data, and wrote the manuscript; V.S. performed protein production and crystallography; M.H. conceived and led the study; K.R., A.S., and V.R.I.K. performed computational studies; H.L., V.S., J.J., R.K., J.K., K.R., and M.H. edited the manuscript; D.L. performed bioinformatics; B.-M.S. performed study design; C.P., I.-S.K., S.G., A.M.O., F.D.F., and J.K. developed, tested and ran the XFEL sample delivery system; H.L., A.S.B., K.D.S., A.Bh., A.Bu., and N.K.S. processed and analyzed XFEL data; F.D.F., R.A.-M., and A.Ba. operated the MFX instrument; H.L., J.J., O.A., C.P., I.-S.K., A.S.B., S.G., K.D.S., A.Bh., P.S.S., A.Bu., P.A., A.M.O., F.D.F., A.Ba., N.K.S., V.K.Y., J.Y., J.K., and M.H. performed the XFEL experiment at LCLS. Competing interests: Authors declare that they have no competing interests. Data and materials availability: Atomic coordinates and structure factors have been deposited in the Protein Data Bank (PDB) with the following codes: catalytically active radical state solved by XFEL, 8bt3; radical-lost ground state, 8bt4. In silico models and output are available at the SciLifeLab Data Repository (49). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse. This research was funded in whole or in part by the Wellcome Trust (grant 210734/Z/ 18/Z). The author will make the Author Accepted Manuscript (AAM) version available under a CC BY public copyright license. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh8160 Materials and Methods Figs. S1 to S5 Table S1 References (50–66) MDAR Reproducibility Checklist Movies S1 and S2 Submitted 17 March 2023; accepted 30 August 2023 10.1126/science.adh8160 Lebrette et al., Science 382, 109–113 (2023) 6 October 2023 5 of 5
10.1126_science.adh8830
RES EARCH HERBIVORY Plant size, latitude, and phylogeny explain within-population variability in herbivory The Herbivory Variability Network*† Interactions between plants and herbivores are central in most ecosystems, but their strength is highly variable. The amount of variability within a system is thought to influence most aspects of plant-herbivore biology, from ecological stability to plant defense evolution. Our understanding of what influences variability, however, is limited by sparse data. We collected standardized surveys of herbivory for 503 plant species at 790 sites across 116° of latitude. With these data, we show that within-population variability in herbivory increases with latitude, decreases with plant size, and is phylogenetically structured. Differences in the magnitude of variability are thus central to how plant-herbivore biology varies across macroscale gradients. We argue that increased focus on interaction variability will advance understanding of patterns of life on Earth. P lant-herbivore interactions, which involve more than half of macroscopic biodiver- sity and 90% of macroscopic biomass (1), are believed to shape macroscale biological patterns and processes, such as plant and herbivore biodiversity gradients, biomass distributions, community structure, species coexistence, and trait evolution (2–4). Biologists have studied the role of herbivory at macroscales by quantifying how the mean herbivore damage level covaries with latitude, biome, functional traits, and phylogeny (5–7). However, macroscale patterns have not always matched expectations. For example, despite the paradigm that herbivore pressure increases toward the equator owing to more-benign envi- ronmental conditions, empirical patterns have been weak or inconsistent (8–10). Similarly, despite the expectation that closely related plant species should face similar pressures from herbivores, phylogenetic signal in mean herbivore damage is often undetectable or restricted to certain groups (5, 11). We suggest that our understanding of macroscale patterns in herbivory can be improved by considering patterns in the magnitude of variability in herbivory rather than only mean interaction strength. Variability is a hallmark of plant-herbivore interactions (12). Within populations, patterns in damage are often highly skewed, with most plant individuals receiving very low levels of damage, and a few plants receiving high levels (13). Although there are limited data on the drivers and consequences of this variability, theory indicates that within-species variation in traits or interactions can be as important as the mean for biological processes ranging from population viability to evolutionary dy- namics (14, 15). For example, spatial variability can stabilize plant-herbivore dynamics by giving plants refuges from overexploitation (16), can increase the importance of competition among herbivores (17), can maintain diversity by fa- cilitating the evolutionary coexistence of al- ternative strategies (18), and can drive disease dynamics by causing superspreading events (19). Variation in damage among plant indi- viduals also indicates the potential pattern of selection by herbivores, which drives plant defense evolution (20). Variability has been hypothesized to favor inducible plant defenses over constitutively expressed defenses—a cen- tral dichotomy in defense evolution (21). Des- pite the central role that variability likely plays in the ecology and evolution of plants and herbivores, macroscale patterns of var- iability remain uncharacterized. In this work, we propose and test three hypotheses for pat- terns in the magnitude of variation in herbi- vore damage among individuals within plant populations. First, we hypothesize that herbivory varia- bility within populations increases with distance from the equator owing to shorter growing seasons and less-stable abiotic conditions at higher latitudes reducing the time available for herbivore foraging. A latitudinal variability gradient could help explain how herbivores have influenced global patterns of plant bio- diversity despite the weak latitudinal gradient in mean herbivory (22, 23). Herbivory may maintain plant diversity at low latitudes not just by being more intense on average but by being a more-consistently important force within plant populations. Second, we hypothe- size that herbivory is more variable among small plants compared with large plants. Large *Corresponding author. Email: william.wetzel@montana.edu †Herbivory Variability Network authors and affiliations are listed in the supplementary materials. low 400 B C high even uneven A i s e c e p s l t n a p f o r e b m u N 200 0 0.0 0.5 Mean herbivory (proportion damage) 1.0 100 50 i s e c e p s l t n a p f o r e b m u N 0 0.0 0.5 Variability in herbivory (Gini coefficient) 1.0 i y r o v b r e h f o n o i t r o p o r p l e v i t a u m u C 1.0 0.5 0.0 Gini coefficient 1.0 0.5 0.0 0.0 1.0 0.5 Cumulative proportion of plants Fig. 1. Mean and variability in plant-herbivore interactions. (A) Histogram of the number of plant species with different mean proportion leaf area damaged by herbivores. (B) Histogram of the Gini coefficient values for all plant species in our dataset. (C) Lorenz curves from all 790 population surveys in our dataset. Each curve shows the cumulative proportion of herbivory across the cumulative proportion of plants, ordered by increasing herbivory, for one plant population. Curves closer to the 1:1 line (gray dashes) indicate more-even distributions. Lorenz curves form the basis for the calculation of the Gini coefficient of inequality, which ranges from 0 (a perfectly even distribution) to 1 (a perfectly uneven distribution). Curves are colored by their Gini coefficient [as in (B)]. Sample sizes are 790 surveys of 503 plant species. Robinson et al., Science 382, 679–683 (2023) 10 November 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E plants, which represent a greater sampling area, should average over small-scale random variation in herbivory, resulting in values closer to the population mean, whereas small plants should be more likely to escape herbivory entirely or be highly damaged by a few events. If supported, this hypothesis would expand our understanding of long-studied differences in defenses between trees and herbs (24), with consistent damage on large plants explaining why trees invest a greater proportion of their biomass in constitutive defenses (25). Third, we hypothesize that variability in herbivory is phylogenetically structured, with more-closely related plants displaying more-similar levels of variability. This pattern, which has been documented for mean herbivory (5), would indicate that variability is influenced by species- level traits and is not simply random, as it has often been treated. To characterize macroscale patterns in population-level mean and variability in her- bivory, 127 research teams in 34 countries used a standardized protocol (26) to sample plants and quantify aboveground herbivore damage for 790 populations of 503 species in 135 fami- lies. This sample comprised more than 50,000 A B i y r o v b r e h n i y t i l i b a i r a V i ) t n e c i f f e o c i i n G ( 1.0 0.5 0.0 0 Gini coefficient 1.0 0.5 0.0 C i y r o v b r e h n a e M ) n o i t r o p o r p ( 1.0 0.5 0.0 20 40 60 Latitude (absolute value) Biome 0 20 40 60 Latitude (absolute value) D i y r o v b r e h n i y t i l i b a i r a V ) t i n e c i f f e o c i i n G ( 0.8 0.7 0.6 0.5 0.4 0.025 0.050 0.075 Mean herbivory (proportion) Tundra Boreal Forests/Taiga Montane Grasslands & Shrublands Temperate Grasslands, Savannas & Shrublands Temperate Broadleaf & Mixed Forests Temperate Conifer Forests Mediterranean Forests, Woodlands & Scrub Tropical & Subtropical Moist Broadleaf Forests Deserts & Xeric Shrublands Tropical & Subtropical Dry Broadleaf Forests Tropical & Subtropical Grasslands, Savannas & Shrublands Fig. 2. Global patterns of variability in herbivory within plant populations. (A) The geographic distribution of our sampling sites, colored by variability in herbivory among individuals within populations (Gini coefficient). Points are slightly jittered for visibility. (B and C) Variability in herbivory increased (B) and mean herbivory decreased (C) with latitude across our sampling extent. Lines show predicted means and 50, 80, and 95% credible intervals from Bayesian phylogenetic beta regressions. (D) The 11 biomes in our study can be characterized by their mean and variability in herbivory. Herbivory variability and mean showed an inverse relationship across biomes [r = −0.67 (−0.94 to −0.08)], but there were also differences in variability between biomes with similar means. Error bars show 50 and 80% credible regions. Sample size is 790 surveys of 503 species. Legend in (D) is ordered by Gini coefficient. Robinson et al., Science 382, 679–683 (2023) 10 November 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E plant individuals distributed across six con- tinents and 116° of latitude. Past macroscale studies that have focused on differences in means typically examined relatively few indi- viduals per population (5). By contrast, we sampled 60 individuals per population, which allowed us to analyze patterns in population- level variability. For each plant individual, we recorded plant size (height for most species, or canopy diameter for prostrate species) and vis- ually estimated the cumulative proportion of leaf tissue damaged by invertebrate and verte- brate herbivores. We quantified the variability in herbivory among individuals within popu- lations using the Gini coefficient—a commonly used scale-invariant metric that ranges from 0 to 1 (perfectly even to perfectly uneven) (27). We tested our hypotheses by quantifying associa- tions between each macroscale factor and the Gini coefficient or mean herbivory using Bayesian phylogenetic beta regressions. Overall, within-population variation in her- bivore damage was very high [mean Gini coefficient = 0.61 (95% confidence interval: 0.40 to 0.78)] (Fig. 1). On average, the most-damaged individual in each plant population lost 34.2% (32.4 to 36.0%) of its leaf area to herbivory, whereas 27.9% (25.9 to 29.9%) of individuals completely or essentially escaped herbivory (<0.5% damage). Half of the damage in each population was concentrated on 11.3% (10.7 to 11.9%) of its individuals on average. The level of variation within populations also varied sig- nificantly across populations and species, with the Gini coefficient ranging from 0.03, an almost perfectly even distribution of damage, to 1.0, a perfectly uneven distribution with all damage on one plant (Fig. 1, B and C). Even though the Gini coefficient normalizes by the mean, it can nevertheless be correlated with it. Mean herbivory and the Gini coefficient were nega- tively correlated, with Gini coefficients being low for the 3.9% of populations with very high (>25%) mean herbivory, whereas populations with lower mean herbivory exhibited the full range of Gini coefficients (r = −0.46) (fig. S1). Geographic patterns of variability We found strong support for the latitudinal variability gradient hypothesis (Fig. 2, A and B). Variation was lowest at the equator [Gini = 0.51 (0.33 to 0.69)] and increased toward 70°N and 70°S (70°N/S) [Gini = 0.70 (0.54 to 0.84); Bayesian coefficient of determination (R2) = 5%; posterior probability (pp) = 1.0; Bayes factor (BF) = 2.0 × 104]. Mean herbivory, by contrast, declined with latitude, from 8.0% (4.1 to 12.3%) at the equator to 2.9% (1.4 to 4.7%) at 70°N/S; this relationship was less predictable than the one for the Gini coefficient (R2 = 2%; pp = 1.0; BF = 2.9 × 104) (Fig. 2C, figs. S2 and S3, and tables S1 to S3). Thus, plants at higher latitudes, with shorter growing seasons and lower tem- peratures (26), receive less herbivory on average, and that herbivory is concentrated on fewer individuals. This result could conceivably be an artifact of the negative mean–Gini coefficient correlation. We therefore repeated our analysis with mean herbivory included as a covariate. The estimated latitudinal variability gradient was still strongly positive, though it was lower in magnitude, with a 20% (6 to 38%) increase in the Gini coefficient from the equator to 70° N/S (R2 = 23%; pp = 1.0; BF = 14.5) (fig. S4). This relationship captured differences among biomes: Higher latitude and higher elevation biomes had higher Gini coefficients and lower mean herbivory (Fig. 2D and fig. S5). Whereas there was a negative correlation between the mean and Gini coefficient among biomes [r = −0.68 (−0.95 to −0.10)], there were also large differences in the Gini coefficient between biomes with similar mean herbivory. This sug- gests that interaction variability could be a fun- damental characteristic differentiating biological systems across macroscales. Debate over the contribution of herbivory to global patterns of plant evolution has been contentious (3, 6, 8, 10, 22, 23). Our data show strong evidence of a meaningful, although noisy, latitudinal decline in mean levels of herbivore damage. They also show that herbivory becomes more variable with increasing latitude. This pattern is consistent with our hypothesis that herbivory influences plant evolution at low latitudes not just by being more intense on average, but also by being more consistently important within a plant population. Theory predicts that the relationship between the strength of antagonistic interactions and the intensity of selection is concave down (saturat- ing) at low mean interaction strengths (28), which means that variability at high latitudes, where mean herbivory is low, should erode selection through nonlinear averaging (14), all else being equal. Our finding is also consistent with the hypothesis that inducible defenses are more common among temperate com- pared with tropical plants (29, 30) because greater variation in herbivory is predicted to select for inducibility (21). In addition to sea- sonality and climate, other mechanisms for the latitudinal variability gradient could in- clude greater predation pressure on herbivores at low latitudes (3) suppressing localized out- breaks and high tropical herbivore diversity and specialization (31) evening out damage patterns across plant individuals. More gen- erally, our results confirm the long-held view that biotic interactions are more consistent in the tropics, perhaps owing to longer grow- ing seasons or greater species diversity and specialization (3). Variability and plant size We also found strong support for the size- mediated variability hypothesis. Populations of larger individuals exhibit less variability in Fig. 3. Plant size shapes variability in herbivory. (A) Variability in herbivory among individuals within populations declines with the average size (height or canopy diameter for prostrate species) of plants in the population (R2 = 13.3%; pp = 1.0; BF = 4.6 × 107; 735 surveys of 472 species). (B) Variability in herbivory, however, is only weakly related to plant growth form (R2 = 2.8%), with woody plants having 10.9% (2.9 to 19.1%) lower Gini coefficients compared with herbaceous species (790 surveys of 503 species). Lines, shaded regions, and large points show predicted means and 50, 80, and 95% credible intervals from phylogenetic Bayesian beta regressions. Each small gray point is one survey. herbivory among individuals. A 2-m increase in mean plant size (from 0.05 to 2.05 m, en- compassing ~90% of our populations) resulted in a 32.7% (20.6 to 44.7%) decrease in the Gini coefficient [from 0.70 (0.54 to 0.85) to 0.47 (0.29 to 0.66); R2 = 13.3%; pp = 1.0; BF = 4.6 × 107] (Fig. 3A and fig. S6). This relationship held even after accounting for the decline in plant size with increasing latitude and differences in plant abundance (which ranged from 2 to 100% cover in our dataset) (tables S4 and S5) (32). Woody species, which averaged 4.1 times as large as herbs in our dataset, had 10.9% (2.9 to 19.1%) lower Gini coefficients compared with herbaceous species [0.56 (0.37 to 0.76) versus 0.63 (0.44 to 0.81); BF = 4.25]. How- ever, the overall variance explained by growth form, including climber and graminoid catego- ries, was low (R2 = 2.8%) (Fig. 3B and fig. S7), which suggests that mean size is a more- important determinant of herbivory patterns than growth form. Mean herbivory, by contrast, Robinson et al., Science 382, 679–683 (2023) 10 November 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fabaceae L a t h y r u s q uin q u e n e r viu M e d i c a g o p o l y m o r p h a Tr i f o l i u m p r a t e n s e t u s alb u s M elilo A s t r a g a l u s a l p i n u s Tr i f o l i u m r e p e n s A s t r a g a u s a u s t r a H e d y s a r u m b o r e a e l i s l l s a s A m A o o r p t u h s a g c e a n n i e s t s o c e d n e s s i i L u p n u s a r g e n t e u s s u l l l s u y h e p r o y b o r p a s s u u n n p p u u L L i i s i s n e x e t s u n p u L i s u s o l i p s u n p u L i s e m biu olo a s u e c i r e s s u n i p u L s i n n e r e p s u n i p u L a d i l l a p a i r a l a t o r C a e c n u j a i r a l a t o r C o r o b ellin vit ia mia r ala r e t o alb r C D s e d A mia a L l S o a n a c e a e a m m e a e c Plantaginaceae Pe In dio P h dig Vicia villo Psorale a m Vicia s Vicia s s ofera hirs elu e olu C Desmodium gangeticum Gly m te ulle Desmodium procumbens Lespedeza pilosa s lu a corylifoliu e ativ cin piu n corylifoliu Desmodium glutinosum n Lespedeza hirta Polygonaceae n uifloru e m uta atu m Desmodium incanum s a x m Rumex japonicus Rumex acetosa Rumex obtusifolius Rumex crispus Rumex sanguineus Coccoloba cereifera Coccoloba uvifera Persicaria virginiana Bistorta bistortoides Bistorta vivipara Hebe salicifolia Digitalis purpurea Veronica officinalis Plantago depressa Plantago hispidula Plantago lanceolata Plantago major Plantago asiatica Stachys grandidentata Hyptis suaveolens Salvia pratensis Salvia sclarea Rosmarinus officinalis Lycopus uniflorus Thymus vulgaris Monarda punctata Monarda fistulosa Nicotiana attenuata Datura wrightii Salpichroa origanifolia Physalis longifolia Physalis heterophylla Solanum nigrum Solanum sp Solanum ptychanthum Solanum dulcamara Solanum lycocarpum Solanum donianum Solanum carolinense Solanum cinereum Solanum incanum Solanum elaeagnifolium Solanum tridynamum Cephalanthus occidentalis Cordiera elliptica Coffea arabica Coprosma lucida Galium circaezans Galium album Galium verum Morinda pubescens Mitchella repens Psychotria aubletiana Palicourea sp Palicourea padifolia Palicourea rigida Landolphia dulcis Apocynum cannabinum Vincetoxicum hirundinaria Calotropis procera Gomphocarpus fructicosus Asclepias speciosa Asclepias cryptoceras Asclepias asperula Asclepias tuberosa Asclepias syriaca Apocynaceae Asclepias curassavica Asclepias verticillata Asclepias incarnata Flourensia thurifera Verbesina encelioides Hulteniella integrifolia Kuhnia eupatorioides Conoclinium coelestinum R u bia c e a e c e b d u R e b e a b R g ata eckia hirta n ciniata a pin a kia trilo errim a eifolia e atibid kia la gin e s gia ferru m min a sto ptalia inte n n e u v stia c r ure a ctiu m nta siu r Pro A e Cir C c e b d u R b d u R a h C n u J s u u c n n a a s o s o a c A a c A a c A a c A a o n n e S nioid s gifolia a gia mis n ata x uiflora ole ple alb e cia im Chamaecrista fasciculata v pis velutin cia lo btusifolia Senna occidentalis e a a albid cia d u a te cia s Montigena novae−zelandiae Senna cumingii Cercis canadensis Bauhinia racemosa Latrobea pinnaculum Mim Proso Mim Rosaceae Potentilla erecta Potentilla simplex Potentilla recta Potentilla gracilis Sibbaldia procumbens Fragaria vesca Agrimonia eupatoria Rubus chamaemorus Rubus fruticosus Sorbus aucuparia Geum rossii Abutilon theophrasti Sphaeralcea coccinea Sida acuta Plagianthus regius Alcea rosea Hampea appendiculata Gossypium hirsutum Talipariti pernambucensis Urena lobata Brachychiton populneus Tilia americana Waltheria indica Melhania ovata Imperata cylindrica Andropogon virginicus Zea mays Zoysia japonica Bouteloua gracilis Elymus hystrix Ammophila breviligulata Pleioblastus chino Pascopyrum smithii M alv a c e a e P o a c e a e Solidago virgaurea Solidago simplex Solidago caesia Solidago speciosa Solidago canadensis Solidago missouriensis Solidago altissima Solidago canadensis.altissima Symphyotrichum novae−angliae Symphyotrichum novi−belgii Symphyotrichum cordifolium Eurybia macrophylla Grindelia squarrosa Heterotheca subaxillaris Erigeron strigosus Erigeron glacialis Erigeron canadensis Baccharis dracunculifolia Baccharis serrulata Kalimeris integrifolia Aster oblongifolius Celmisia discolor Celmisia spectabilis Olearia paniculata Anthemis galilae Achillea ptarmica Leucanthemum vulgare Artemisia ludoviciana Artemisia mongolica Senecio viscosus Senecio madagascariensis Senecio elegans Senecio crassulus Senecio brunonianus Jessea multivenia Emilia praetermissa Jacobaea vulgaris Tussilago farfara Helianthella quinquenervis Encelia canescens Helianthus annuus H Echin elia E c nth B hin M als ace ethia m us occid ela a C c a eratu a p e m h nth a a r ratu orhiz o urp m era s ollis n m fa g ola m c ure ntalis u a s stifolia e c a stigiatu a a n Asteraceae gittata n a o d e d n o s r ata o n y z oid e s A g e W y A g m e a e t n a g i g a i n o n r e V i i n i w d l a b a i n o n r e V a tit r a tip a n pin zia o n n u M s u i b u d n o g o p o g a Tr a l l e s o l i p m u i c a r e i H a c i r i b i s a c u t c a L l i l m u r o h p o t a r e c m u c a x a r a T e m a u n c i c i f f o m u c a x a r a T o g n o m m u c a x a r a T i s u m s s i t l a s u a b a N l P u B i d e n s r e p t a n s A n t e n n a r i a n e g e c t a l i c a r i a d y s e n t e r i c a l H H y y p p o o c c h h a a e e r r i i s s r c a h d c a t a e n s s i l l i i B i d e n s f r o n d o s a A r n i c a c o r d i f o l i a Tr i d a x p r o c u m b e n s A A g g e e r a r a tin tin a a h altis a v a n sim e n sis a Mean herbivory (proportion) 0.0 0.5 1.0 Variability in herbivory (Gini coefficient) 0.0 0.5 1.0 Fig. 4. Phylogenetic patterns of mean and variability in herbivory. Variability in herbivory among plants within populations (Gini coefficient) show greater phylogenetic signal [Pagel’s l = 0.51 (0.45 to 0.52); P < 0.001] compared with mean herbivory levels [Pagel’s l = 0.07 (0.06 to 0.08); P > 0.1]. For clarity, this tree includes only the 240 species from the 11 best-represented plant families (≥8 species per family). Our analyses included all 503 species in the dataset (see fig. S10 for the full tree). was unrelated to mean size or growth form (figs. S8 and S9). We posit that lower among-individual vari- ability in herbivory on large plants results from the law of large numbers, which says that processes that involve more random events produce values closer to the overall mean. In other words, large plants, which have a greater number of potential herbivory events, average over within-plant variability and receive values closer to the population mean on average. Small plants, by contrast, are more likely to escape herbivory entirely or be severely damaged by a few events, which results in high variability. A key implication of this phenomenon is that Robinson et al., Science 382, 679–683 (2023) 10 November 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E larger species (and larger stages within species) should experience greater selection for high concentrations of constitutive defenses or tolerance. Smaller species (and stages), by con- trast, should experience greater selection for inducible defenses and low concentrations of metabolically cheap toxins to save resources in the absence of herbivory and repel herbivores when encountered. This dichotomy in defense evolution has been the focus of decades of research on differences in defenses between trees and herbs (24) and across ontogenetic stages (33). Whereas previous work has in- voked complex biological explanations for these differences—such as how apparent plants are to herbivores (24)—our results suggest that patterns are more parsimoniously ex- plained by the statistical consequences of mean plant size. Phylogenetic patterns of variability Finally, we tested the hypothesis that variability in herbivory is phylogenetically structured. The Gini coefficient exhibited significant phyloge- netic signal [Pagel’s l = 0.51 (0.45 to 0.52); P < 0.001], indicating that more-closely re- lated species display more-similar variability levels (Fig. 4 and fig. S10). Mean herbivory, by contrast, did not show meaningful phyloge- netic signal [l = 0.07 (0.06 to 0.08); P = 1.0]. These results were robust to tree topology and species sampling (supplementary materials). Our findings suggest that the mean damage level across species changes relatively rapidly in response to evolutionarily labile plant traits, whereas the variability is more strongly de- termined by traits that are phylogenetically conserved. Traits thought to influence the amount of herbivore damage, such as chemi- cal defenses, diverge as plants escape their herbivores by evolving novel defenses (2, 34), whereas characteristics such as geographic lo- cation and plant size, which we find relate to variability, tend to be less labile. High varia- bility in some families (e.g., Apocynaceae and Plantaginaceae) invites further investigation and could help reveal drivers of these conserved patterns. To examine macroevolutionary pat- terns, we fit Brownian motion and Ornstein- Uhlenbeck models to test for differences in rates of evolution and the strength of stabiliz- ing selection. The best-fitting models included optima for variability and mean herbivory in tropical versus temperate systems and woody versus herbaceous growth forms (tables S6 and S7), which indicates that the evolution of variability in herbivory seems to be driven by conserved plant traits and is therefore a biologically informative feature rather than random noise. Conclusions The assumption that plant-herbivore interac- tions are highly variable has long dominated ecology and evolution, with foundational works on so-called variable plants and herbivores (12) and theory exploring the consequences of variable herbivory (21). Our data confirm this assumption but also reveal a pattern that had not been previously documented: strong differentiation across systems in the level of variability itself. Variation in herbivory co- varied with factors central to the ecology and evolution of plant-herbivore interactions, such as latitude, biome, plant size, and phylogeny. These macroscale patterns were often stronger than patterns for mean herbivory levels. This suggests that the level of variability could be important for driving differences in plant- herbivore biology around the planet, between species with different traits and across phylo- geny. Although the importance of varia- bility in interactions has been recognized by a few fields, such as epidemiology (19), the central role of interaction variability in shap- ing macroscale patterns of life on Earth has been underappreciated. Our global dataset is evidence for the ubiquity and predictabil- ity of variability in one biotic interaction and highlights the promise of further explorations of the causes and consequences of interaction variability. RE FERENCES AND NOTES 1. Y. M. Bar-On, R. Phillips, R. Milo, Proc. Natl. Acad. Sci. U.S.A. 115, 6506–6511 (2018). 2. P. R. Ehrlich, P. H. Raven, Evolution 18, 586–608 (1964). 3. D. W. Schemske, G. G. Mittelbach, H. V. Cornell, J. M. Sobel, K. Roy, Annu. Rev. Ecol. Evol. Syst. 40, 245–269 (2009). 4. O. J. Schmitz, Annu. Rev. Ecol. Evol. Syst. 39, 133–152 (2008). 5. M. M. Turcotte, T. J. Davies, C. J. M. Thomsen, M. T. J. Johnson, Proc. R. Soc. B 281, 20140555 (2014). 6. A. T. Moles et al., New Phytol. 191, 777–788 (2011). 7. S. Rasmann, A. A. Agrawal, Ecol. Lett. 14, 476–483 (2011). 8. A. T. Moles, S. P. Bonser, A. G. B. Poore, I. R. Wallis, W. J. Foley, Funct. Ecol. 25, 380–388 (2011). 9. D. Salazar, R. J. Marquis, Proc. Natl. Acad. Sci. U.S.A. 109, 12616–12620 (2012). 10. J. Y. Lim, P. V. A. Fine, G. G. Mittelbach, Glob. Ecol. Biogeogr. 24, 1106–1112 (2015). 11. X. Moreira et al., Phytochemistry 153, 64–73 (2018). 12. R. F. Denno, M. S. McClure, in Variable Plants and Herbivores in Natural and Managed Systems, R. F. Denno, M. S. McClure, Eds. (Academic Press, 1983), pp. 1–12. 13. W. C. Wetzel, B. D. Inouye, P. G. Hahn, S. R. Whitehead, N. Underwood, Annu. Rev. Ecol. Evol. Syst. 10.1146/annurev- ecolsys-102221-045015 (2023). 14. D. I. Bolnick et al., Trends Ecol. Evol. 26, 183–192 (2011). 15. C. Violle et al., Trends Ecol. Evol. 27, 244–252 (2012). 16. R. M. Anderson, R. M. May, J. Anim. Ecol. 47, 219–247 (1978). 18. A. A. Agrawal, J. A. Lau, P. A. Hambäck, Q. Rev. Biol. 81, 349–376 (2006). 19. J. O. Lloyd-Smith, S. J. Schreiber, P. E. Kopp, W. M. Getz, Nature 438, 355–359 (2005). 20. A. A. Agrawal, A. P. Hastings, M. T. J. Johnson, J. L. Maron, J.-P. Salminen, Science 338, 113–116 (2012). 21. F. Adler, R. Karban, Am. Nat. 144, 813–832 (1994). 22. A. T. Moles, J. Ollerton, Biotropica 48, 141–145 (2016). 23. D. N. Anstett, K. A. Nunes, C. Baskett, P. M. Kotanen, Trends Ecol. Evol. 31, 789–802 (2016). 24. P. Feeny, in Biochemical Interaction Between Plants and Insects, J. W. Wallace, R. L. Mansell, Eds. (Springer, 1976), pp. 1–40. 25. A. M. Smilanich, R. M. Fincher, L. A. Dyer, New Phytol. 210, 1044–1057 (2016). 26. See the supplementary materials and methods. 27. J. L. Gastwirth, Rev. Econ. Stat. 54, 306–316 (1972). 28. C. W. Benkman, Ecol. Lett. 16, 1054–1060 (2013). 29. R. J. Bixenmann, P. D. Coley, A. Weinhold, T. A. Kursar, Ecol. Evol. 6, 6037–6049 (2016). 30. P. D. Coley, M.-J. Endara, T. A. Kursar, Oecologia 187, 361–376 (2018). 31. L. A. Dyer et al., Nature 448, 696–699 (2007). 32. A. T. Moles et al., J. Ecol. 97, 923–932 (2009). 33. K. Boge, R. J. Marquis, Trends Ecol. Evol. 20, 441–448 (2005). 34. T. A. Kursar et al., Proc. Natl. Acad. Sci. U.S.A. 106, 18073–18078 (2009). 35. W. Wetzel et al., Data for: Plant size, latitude, and phylogeny explain within-population variability in herbivory, data set, Dryad (2023); https://doi.org/10.5061/dryad. 44j0zpckm. 36. W. C. Wetzel, HerbVar-Network/HV-Large-Patterns-MS-public: v1.0.0, Zenodo (2023); https://doi.org/10.5281/zenodo.8133118. AC KNOWLED GME NTS The authors thank A. Agrawal, J. Thaler, J. Conner, J. Lill, M. Weiss, N. Sanders, and M. Meek for helpful discussions and comments on different versions of the analyses and manuscripts. Any use of trade, firm, or product names is for descriptive purposes only and does not imply endorsement by the US government. The findings and conclusions in this publication are those of the authors and should not be construed to represent any official US Department of Agriculture (USDA) or US government determination or policy. Funding: The authors acknowledge funding for central project coordination from NSF Research Coordination Network grant DEB-2203582; the Ecology, Evolution, and Behavior Program at Michigan State University; and AgBioResearch at Michigan State University. Site-specific funding is listed in the supplementary materials. Author contributions: This project was conceptualized by W. C. Wetzel and N. Underwood. It was coordinated by W. C. Wetzel, M. L. Robinson, P. G. Hahn, B. D. Inouye, N. Underwood, S. R. Whitehead, K. C. Abbott, E. M. Bruna, N. I. Cacho, and L. A. Dyer. Other author contributions are listed in the supplementary materials. Competing interests: The authors declare that they have no competing interests. Data and materials availability: The dataset generated and analyzed in this study is available at Dryad (35). Our code is archived at Zenodo (36). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https:// www.science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh8830 Herbivory Variability Network Authors Materials and Methods Supplementary Text Figs. S1 to S10 Tables S1 to S7 References (37–58) MDAR Reproducibility Checklist 17. R. F. Denno, M. S. McClure, J. R. Ott, Annu. Rev. Entomol. 40, 297–331 (1995). Submitted 26 March 2023; accepted 27 September 2023 10.1126/science.adh8830 Robinson et al., Science 382, 679–683 (2023) 10 November 2023 5 of 5
10.1126_science.adh8190
RES EARCH BIOPHYSICS MyoD-family inhibitor proteins act as auxiliary subunits of Piezo channels Zijing Zhou1,2†, Xiaonuo Ma3,4†, Yiechang Lin5, Delfine Cheng1,2, Navid Bavi6, Genevieve A. Secker7, Jinyuan Vero Li1,2, Vaibhao Janbandhu1,2, Drew L. Sutton7,8, Hamish S. Scott7,8,9, Mingxi Yao10, Richard P. Harvey1,2,11, Natasha L. Harvey7,8, Ben Corry5, Yixiao Zhang3,4*, Charles D. Cox1,12* Piezo channels are critical cellular sensors of mechanical forces. Despite their large size, ubiquitous expression, and irreplaceable roles in an ever-growing list of physiological processes, few Piezo channel– binding proteins have emerged. In this work, we found that MyoD (myoblast determination)–family inhibitor proteins (MDFIC and MDFI) are PIEZO1/2 interacting partners. These transcriptional regulators bind to PIEZO1/2 channels, regulating channel inactivation. Using single-particle cryogenic electron microscopy, we mapped the interaction site in MDFIC to a lipidated, C-terminal helix that inserts laterally into the PIEZO1 pore module. These Piezo-interacting proteins fit all the criteria for auxiliary subunits, contribute to explaining the vastly different gating kinetics of endogenous Piezo channels observed in many cell types, and elucidate mechanisms potentially involved in human lymphatic vascular disease. T o decode mechanical cues, cells are en- dowed with a palette of molecular force sensors. Among these sensors, Piezo ion channels (1) have emerged as critical force sensors that participate in determining how cells sense their physical environment. Piezo channels assemble as trimers that pos- sess all the structural requirements for mecha- nosensitivity (2, 3). However, native PIEZO1 channels can display nonuniform subcellu- lar localization (4–7) and exhibit different gating kinetics—principally, slower inactiva- tion (1, 8–13)—in many cell types when com- pared with heterologous expression systems. These observations could be explained by dif- ferences in lipid composition (14, 15), curvature- dependent sorting (5, 7), or protein-protein interactions. Many ion channels interact with auxiliary subunits (16, 17) to modify their cellular location and gating properties. Ion- 1Victor Chang Cardiac Research Institute, Sydney, NSW 2010, Australia. 2School of Clinical Medicine, Faculty of Medicine and Health, University of New South Wales, Sydney, NSW 2052, Australia. 3Interdisciplinary Research Center on Biology and Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Shanghai 200032, China. 4State Key Laboratory of Chemical Biology, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, Shanghai 200032, China. 5Research School of Biology, Australian National University, Acton, ACT 2601, Australia. 6Department of Biochemistry and Molecular Biophysics, University of Chicago, Chicago, IL 60637, USA. 7Centre for Cancer Biology, University of South Australia and SA Pathology, Adelaide, SA 5001, Australia. 8Adelaide Medical School, University of Adelaide, Adelaide, SA 5005 Australia. 9Department of Genetics and Molecular Pathology, SA Pathology, Adelaide, SA 5000, Australia. 10Department of Biomedical Engineering, Southern University of Science and Technology, Shenzhen 518055, China. 11School of Biotechnology and Biomolecular Science, University of New South Wales Sydney, Kensington, NSW 2052, Australia. 12School of Biomedical Sciences, Faculty of Medicine and Health, University of New South Wales Sydney, Kensington, NSW 2052, Australia. *Corresponding authors. Email: c.cox@victorchang.edu.au (C.D.C.); yzhang@mail.sioc.ac.cn (Y.Z.). †These authors contributed equally to this work. channel auxiliary subunits are commonly defined by four criteria: (i) they are non–pore-forming subunits, (ii) they have a direct and stable in- teraction with the pore-forming subunit, (iii) they modulate channel properties in heterol- ogous systems, and (iv) they regulate endoge- nous channel activity in native cells (17). Despite substantial effort, to our knowledge, no Piezo channel–binding partners or auxiliary subunits have emerged that fit these criteria. Identification of Piezo channel–binding proteins To identify binding partners for Piezo channels we used two complementary affinity-capture mass spectrometry (AC-MS) strategies in con- junction with two CRISPR-Cas9–edited, PIEZO1- expressing, human dermal fibroblast (hDF) lines (Fig. 1, A to F, and fig. S1). The reason we chose fibroblasts lies in the previously reported slower inactivation kinetics of PIEZO1 in this cell type (10, 12, 13). Our pipeline consisted of three comparator groups: (i) hDF with PIEZO1 ablated with CRISPR/cas9 (that were com- pared with wild-type cells in which PIEZO1 was enriched through a conventional anti- PIEZO1 antibody strategy), (ii) hDF with a HaloTag added into the endogenous PIEZO1 loci (P1-Halo) where PIEZO1 was enriched with HaloTrap resin, and (iii) primary human car- diac fibroblasts (hCF) in which PIEZO1 was enriched through a conventional anti-PIEZO1 antibody strategy. We then stringently ana- lyzed the resulting MS data to identify PIEZO1- interacting proteins present in all three groups that were not present in any of their respective negative controls. Using these criteria, we only identified two proteins, the first of which was PIEZO1 (Fig. 1E). This provided strong val- idation for both affinity-capture strategies. The second protein identified was the sparsely studied transcriptional regulator MyoD (myo- blast determination) family–inhibitor domain- containing protein (MDFIC) (18, 19). MS provided 39 ± 14% coverage of MDFIC protein averaged over the three groups (Fig. 1F). Using RNA expression data, we identified many cell types in addition to fibroblasts that coex- press Piezo channels and the MyoD-family in- hibitor proteins, MDFIC or MDFI (fig. S2, A and B) (20). We then validated the protein-protein interaction by expressing N-terminally hemag- glutinin (HA)–tagged MDFIC together with PIEZO1 in human embryonic kidney 293 cells with SV40 large-T antigen (HEK293T). Using a coimmunoprecipitation (co-IP) assay, we could reciprocally capture PIEZO1 with MDFIC and found that the complex was present under mechanical (shear stress) or chemical [10 mM Yoda-1 (21)] activation of PIEZO1 (fig. S2C), which indicated the stability of the interac- tion. We also confirmed that PIEZO1 interacted with the closely related MDFI (fig. S2D) (22). PIEZO2 has high sequence similarity with PIEZO1, so we next tested whether MDFIC se- lectively interacted with PIEZO1/2 channels using native gels. We identified MDFIC at the size of the respective PIEZO1/2 trimers but not in oligomeric complexes of other structurally unrelated channels such as TRPM4 (tetramer) or TREK-1 (dimer) (fig. S2E). In doing so, we found that PIEZO1 enhanced the protein amounts of both MDFIC and MDFI by greater than three- to fourfold (fig. S2, C to E). To determine the specificity of this effect, we coexpressed MDFIC with PIEZO1 and PIEZO2 and compared the amount of MDFIC when expressed alongside other unrelated ion channels. Coexpression of MDFIC with green fluorescent protein (GFP), TRPM4, TRPV4, or TREK-1 did not enhance the amount of MDFIC (fig. S3, A and B). To understand the mechanism, we used a cyclo- heximide chase assay to observe the degrada- tion time course of MDFIC. In cells treated for 4 to 8 hours with the protein synthesis inhib- itor cycloheximide, we observed using Western blotting that the protein amount of MDFIC fell much more rapidly in the absence of PIEZO1 (fig. S3, C and D). This suggests that interacting with PIEZO1 decreased MDFIC turnover. MDFIC binds the PIEZO1 pore module To provide molecular details of the interaction, we coexpressed mouse PIEZO1 (mPIEZO1) with N-terminally FLAG-tagged mouse MDFIC and purified the complex using FLAG resin. Using single-particle cryogenic electron microscopy (cryo-EM), we determined the structure of the PIEZO1-MDFIC complex at an overall resolu- tion of 3.66 Å (Fig. 2, A and B, and fig. S4). We resolved the C-terminal 21 amino acids of MDFIC (Fig. 2, A to D), whereas the N-terminal portion displayed little or no density, presumably ow- ing to local dynamics. The resolved region of MDFIC (residues 225 to 246) consists of an amphipathic helix that sits parallel to the membrane at the membrane interface (Fig. 2C). This helix inserts laterally into the PIEZO1 Zhou et al., Science 381, 799–804 (2023) 18 August 2023 1 of 6 A D RES EARCH | R E S E A R C H A R T I C L E Fig. 1. AC-MS identifies a newly characterized family of Piezo-channel binding partners. Groups for AC-MS consisted of: (A) WT and PIEZO1-edited (knockout, KO) human dermal fibroblasts (hDF; n = 2), (B) WT and PIEZO1-HaloTag (P1-Halo) hDF (n = 2), and (C) primary human cardiac fibroblasts (hCF; n = 2). (D) Affinity-captured protein lysates were run on SDS–polyacrylamide gel electrophoresis (SDS-PAGE) gels and sectioned into quadrants (red dashed lines). Each quadrant was subjected to in-gel protein digestion, peptide extraction and liquid chromatography (LC), and MS, in that order. (E) Venn diagram illustrating proteins identified in each experimental group that had ≥2 distinct peptides that were absent from negative control replicates. (F) Two proteins were identified in all positive control replicates: PIEZO1 and MDFIC. Alignment of distinct MDFIC peptides identified with MS. Group 1 B Group 2 C P1-KO hDF WT hDF WT hDF P1-Halo hDF Group 3 hCF Negative control + Piezo1 antibody + Halo Trap Agarose beads + Mouse IgG Piezo1 antibody Antibody affinity capture HaloTag affinity capture Antibody affinity capture KO WT PIEZO1 E hCF F 2. MDFIC 1. PIEZO1 2. MDFIC 6 2 0 0 3 0 9 In-gel protein digestion + peptide extraction 260 160 80 60 50 40 30 20 15 10 LC-MS/MS Analysis Gene Annotation hDF WT vs KO hDF P1-Halo vs WT 0 peptides in control >2 unique peptides C outer leaflet cap inner leaflet MDFIC IH OH α2 α1 α3 G L2127 OH OH I2195 α1 F2120 E229 K2184 D232 S230 E235 V2187 K2183 E239 S231 L234 H2116 F2120 ~90° H2116 ~90° V2187 K2184 K2183 R2169 α3 T2171 W2140 S2148 L2149 S2150 α3 M238 S247 Q2123 α1 F2484 F245 B 90° F A MDFIC E blade IH F2485 I2195 OH M2191 V2187 F2484 C244 F245 I243 C241 C240 L2127 Q2123 α1 C237 I236 M238 F2120 H2116 C233 L234 I2186 S230 D H OH α1 IH α1 IH 90° OH OH IH α1 MDFIC α3 N2161 F2494 α1 I2164 P246 M2493 E2495 S247 F245 α3 S2150 L2149 OH Fig. 2. Structural elucidation of the PIEZO1-MDFIC complex. (A and B) Cryo-EM density maps of the mouse PIEZO1-MDFIC complex at 3.66-Å nominal resolution viewed from the top (A) and side (B), with the resolved MDFIC region colored green. (C) The distal C-terminal of MDFIC resides parallel to the bilayer between the anchor domain (a1 to a3) and the outer helix (OH). The cytoplasmic constriction residues Met2493 and Phe2494 are shown in cyan. (D) The C-terminal region of MDFIC penetrates deep into the pore module of PIEZO1, approaching the inner helix (IH). (E to H) The interactions between PIEZO1 and MDFIC from the (E) membrane-facing view, (F) lateral view, and (G and H) cytoplasmic-facing view. Variants linked to lymphatic malformations (mPIEZO1 V2187 and mMDFIC F245) are labeled in bold. pore module, nestling between the anchor do- main and the outer helix of PIEZO1 (Fig. 2, C and D), making contacts with His2116, Phe2120, and Gln2123 from the a1 helix of the anchor domain; Val2187, Met2191, and Ile2195 from the outer helix; and Phe2484 and Phe2485 at the base of the inner helix that lines the PIEZO1 pore (Fig. 2E). The amphipathic helix of MDFIC con- sists of a sequence of five cysteines that point toward the bilayer interior (Fig. 2, E and F) and a sequence of four negatively charged residues that point toward the solvent, forming salt bridges with multiple lysine residues (Fig. 2, F and G). The MDFIC C terminus penetrates far enough to come close to the cytoplasmic con- striction formed by Met2493 and Phe2494 (Fig. 2H) and residues critical for voltage-dependent inactivation, Lys2479 and Arg2482 (23, 24). Despite its central location, MDFIC binding did not influence the closed structure of the PIEZO1 pore module (fig. S4I). Because both PIEZO1 and MDFIC are es- sential for lymphatic development in mice and humans (19, 25, 26), we investigated using the ClinVar database whether any disease-causing mutations were located within this binding interface. We found mutations in both PIEZO1 (human V2171f; Fig. 2, E and G) and MDFIC (human F244L; Fig. 2, E and H) associated with human lymphatic disease. MDFIC and MDFI regulate Piezo gating Given the location of MDFIC binding, we next tested whether human MDFIC and MDFI- modified human PIEZO1 (hPIEZO1) gating in HEK293T cells. MDFIC and MDFI expressed Zhou et al., Science 381, 799–804 (2023) 18 August 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. MyoD-family inhibitor proteins regulate PIEZO1 and PIEZO2 channel gating. (A to E) Representative cell-attached patch-clamp recordings from hPIEZO1 (A) control and hPIEZO1 in the presence of (B) hMDFIC, (C) hMDFI, (D) the conserved C terminus of MDFIC (hMDFIC C-81), and (E) with MDFIC lacking its C-terminal 20 amino acids (hMDFIC DC20) all at a holding potential of –65 mV. (F to H) Quantification of (F) peak currents per patch, (G) percent current remaining, and (H) normalized current 1 s after pressure release (normalized Ipost) for replicates of cell-attached recordings shown in (A) to (E). (I to L) Representative cell-attached recordings from (I) hPIEZO2 control and hPIEZO2 in the presence of hMDFIC and hMDFI, and [(J) to (L)] quantification of replicates. (M to O) Representative cell-attached patch-clamp recordings from mouse cardiac fibroblasts isolated at E16.5 from (M) WT, (N) heterozygous, and (O) homozygous MdficM131fs* mice. (P to R) Quantification of (P) peak current per patch, (Q) percent current remaining, and (R) normalized Ipost for replicates of cell-attached recordings shown in [(M) to (O)] All data are displayed as mean ± SEM or as maximum-to-minimum box-and-whiskers plot. P values are noted above the plots and were determined with one-way analysis of variant (ANOVA) and either Dunnett’s or Tukey’s multiple comparison test. ns, not significant. Zhou et al., Science 381, 799–804 (2023) 18 August 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E alone did not generate stretch-activated cur- rents (fig. S5, A to D). Compared with hPIEZO1 alone, coexpression with MDFIC resulted in a mild right shift in the pressure response curve [from 14.5 ± 3.2 mmHg (n = 9 cells) to 23.9 ± 3.9 mmHg (n = 6)] (fig. S5, E to G), a marked increase in the peak stretch-evoked currents, a substantial slowing of channel inactivation, and continued channel gating even after the pressure was released (Fig. 3, A to H, and fig. S5). We quantified the latter of these effects using the current that remained 1 s after ap- plication of stretch (Ipost). Neither MDFIC nor MDFI influenced mRNA TMEM150c, demon- strating that these inactivation effects were in- dependent of TMEM150c (fig. S5N) (27). MDFIC did not influence PIEZO1 protein amounts in either HEK293T or LNCaP cells transfected with MDFIC (fig. S6, A to D) but did increase PIEZO1 single-channel conductance from 26 ± 3pS (n = 4) to 48 ± 8 pS (n = 4) (fig. S6, E to G). Thus, the increase in stretch-activated cur- rents in the presence of MDFIC is likely driven by changes in conductance and its strong -+ - + + -- + + + + + + + + + D +hPIEZO1-HaloTag HA-Tag WGA -HA- hMDFIC +HA- hMDFIC <0.0001 <0.0001 G 125 100 ) x a m I % ( H <0.0001 t s o p t n e r r u c g n n a m e R i i 75 50 25 0 <0.0001 1.0 0.8 0.6 0.4 0.2 0.0 I d e z i l a m r o N K (9) 5C-S (9) (10) 5C-A hMDFIC hMDFIC hMDFIC (9) 5C-S (9) (10) 5C-A hMDFIC hMDFIC hMDFIC A B E C241 C237 C244 C240 C233 S230 C HAM PIEZO1 HA-hMDFIC mPEG-Mal10kDa HAM - + - + 10000 F hMDFIC hMDFIC 5C-A hMDFIC 5C-S 100 pA 20 pA 20 pA ) A p ( t n e r r u c k a e P c/a 0 500 ms 500 ms 500 ms -60 mmHg -60 mmHg -60 mmHg 1000 0.004 0.002 100 10 (9) 5C-S (9) (10) 5C-A hMDFIC hMDFIC hMDFIC I PIEZO1 pore MDFIC (lipidated) J MDFIC POPC bilayer MDFIC (unmodified) ) % ( y c n a p u c c O t c a t n o C 100 80 60 40 20 Lipidated 5C-S Unmodified Inner helix 0 L2121 G2124 V2128 P2129 L2197 F2198 A2201 A2201 I2202 V2471 L2475 V2477 F2480 V2481 L2475 F2480 V2481 MDFIC (lipidated) Fig. 4. C-terminal lipidation of MDFIC underlies regulation of PIEZO1 channel gating. (A) Cryo-EM density map showing extension of density on cysteine residues within the C terminus of MDFIC. (B) Representative immunoblot (IB) with Avidin (top) or anti-HA (bottom) from acyl biotin exchange showing lysate ± hydroxylamine (HAM). Blots show a band at the correct size for HA-tagged MDFIC, but the biotinylated MDFIC can only be seen in the HAM-treated group, which indicates palmitoylated- MDFIC (palm-MDFIC). (C) Mass-tagging of MDFIC with methoxy-polyethylene glycol (mPEGylated) maleimide reveals at least three palmitoylation sites labeled 1 to 3 (* represents a nonspecific band in all groups). (D) Membrane-localized HA-tagged MDFIC in PIEZO1-HaloTag expressing HEK293T cells and the wheat germ agglutinin (WGA)– delineated membrane. (E) Representative cell-attached recordings of hPIEZO1 in the presence of WT MDFIC and mutation of the five C-terminal cysteine residues to either alanine (5C-A) or serine (5C-S), which abolishes the regulatory effects of MDFIC. (F to H) Quantification of (F) peak currents per patch, (G) percent current remaining, and (H) normalized Ipost for replicate recordings shown in (E). (I) All-atom molecular dynamic simulations of the mPIEZO1 pore module in complex with unmodified (green), 5C-S mutant (purple), and lipidated MDFIC (pink) C-termini. (J and K) Contact occupancy for MDFIC monomers with PIEZO1 residues from replicate simulations and a snapshot of the location of three inner helix residues important for inactivation that interact with lipidated MDFIC for prolonged periods. P values are noted above the plots and were determined with one-way ANOVA and Dunnett’s multiple comparison test. ns, not significant. Zhou et al., Science 381, 799–804 (2023) 18 August 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E effect on inactivation. The slow “closure” after pressure release—identified in single-channel re- cordings (fig. S6, H to J) and from macrocurrents— was more pronounced at hyperpolarizing volt- ages (fig. S7, A and B), which suggests an effect on the structural transition–governing voltage- dependent inactivation (23, 24). Consistent with this, the channel could be rapidly closed by flipping the voltage to depolarizing potentials (fig. S7, C and D). We identified slower inactiv- ation in both cell-attached and whole-cell modes (fig. S8). Additionally, MDFIC expression did not influence the function of mechanosensitive TREK-1 channels, ruling out nonspecific effects (fig. S9). The C termini of MDFIC and MDFI are highly homologous, whereas the N termini bear little resemblance and have no known function (fig. S10A); therefore, we asked whether the MyoD- inhibitor domain of MDFIC (residues 165 to 246, which we name C-81) expressed alone could modify PIEZO1 function. Despite being unsta- ble, this domain reduced channel inactivation (Fig. 3, D and G, and fig. S10B). Moreover, the amount of C-81 protein was dramatically in- creased when coexpressed with PIEZO1 (fig. S10, C and D). Conversely, truncation of the C- terminal 20 residues (DC20) prevented MDFIC from regulating PIEZO1 gating (Fig. 3, E to G). This truncation reduced the interaction with PIEZO1, and MDFIC DC20 protein amounts were only marginally affected by PIEZO1 (fig. S10, D and E). Thus, functional regulation of the PIEZO1-MDFIC complex depends on the conserved C terminus of MDFIC. Given the homology of PIEZO1 and PIEZO2 in the MDFIC-binding region, we additionally showed that the same regulation occurs for both hPIEZO2 and mouse Piezo orthologs (Fig. 3, I to L, and fig. S11). Moreover, the ex- pression of MDFIC and MDFI correlated with the native inactivating phenotype of PIEZO1 in multiple human and mouse cell lines (fig. S12). This included HEK293T, in which fast PIEZO1 inactivation was seen, and prostate cancer cell lines (LNCaP and DU145), in which differential stretch-activated kinetics were pre- viously reported (9). HEK293T and LNCaP have almost undetectable amounts of MDFIC and MDFI and display fast inactivation, whereas DU145 has notable amounts of MDFI at the mRNA level and inactivates slowly (fig. S12, A to F). Overexpression of MDFIC in LNCaP cells markedly slowed inactivation of native PIEZO1 channels (fig. S12, B to F). In mouse embryonic fibroblasts and C2C12 myoblasts, we detected MDFIC, and both cell types displayed slow PIEZO1 inactivation indicated by larger cur- rent remaining at the end of pressure pulses (fig. S12, G to I). By contrast, in Neuro2A (N2A) cells, PIEZO1 channels exhibited rapid inactivation with undetectable expression of MDFIC and MDFI. We next asked whether genetic loss of MDFIC could modify the kinetics of native PIEZO1 cur- rents. To investigate this, we used a mouse mod- el of a complex lymphatic anomaly known as central conducting lymphatic anomaly, bear- ing a truncated variant of MDFIC that lacked the complete conserved C-terminal but retained the N-terminal region (19). Our assessment of PIEZO1 activity in embryonic cardiac fibro- blasts isolated from embryonic day 16.5 (E16.5) wild-type (WT/WT), heterozygous (WT/M131fs*), and homozygous mice (M131fs*/M131fs*) re- vealed that fibroblasts harboring C-terminally truncated MDFIC had smaller PIEZO1 cur- rents that inactivated more rapidly (Fig. 3, M to R). This effect on inactivation was pheno- copied by silencing MDFIC in fibroblasts with small interfering RNA (siRNA) (fig. S13, D to J). PIEZO1 regulation requires MDFIC lipidation Cryo-EM maps of MDFIC revealed extra den- sities on Cys233, Cys237, Cys240, Cys241, and Cys244 (Fig. 4A). Given that this helix is situated at the membrane interface and contains motifs for cysteine lipidation (28), we hypothesized that these extra densities resulted from lipidation, which is the covalent addition of acyl chains to amino acids (29). A second cryo-EM complex of the MDFIC mutant Cys240Ala (3.68 Å) spe- cifically lacked the extra density at Cys240, which is consistent with posttranslational modification (fig. S14). Acyl biotin exchange confirmed that MDFIC was palmitoylated (Fig. 4B). Mass tagging with polyethylene gycol–modified (PEGylated) maleimide (increases mass per palmitoylated site) showed that there were at least three sites for lipidation on MDFIC (Fig. 4C), that two were in the distal C terminus (fig. S15), and that like most lipidated proteins, MDFIC associated with the plasma membrane (Fig. 4D). To check whether MDFIC lipidation is func- tionally relevant, we mutated all five C-terminal cysteines to alanine or serine. MDFIC with Cys233/237/240/241/244Ala (5C-A) or Cys233/237/ 240/241/244Ser (5C-S) bound to PIEZO1, but nei- ther mutant influenced PIEZO1 activity, which suggests that the lipidation of the MDFIC C terminus is critical for PIEZO1 regulation (Fig. 4, E to H, and fig. S15C). To probe the role of lipidation, we used all- atom molecular-dynamics simulations of the PIEZO1 pore module (residues 1956 to 2547) in complex with the MDFIC C terminus (Fig. 4, I to K, and fig. S16). Simulations consisted of three MDFIC monomers: one unmodified, one lipidated at all five cysteine residues in the helix, and one where each cysteine was mutated to serine (5C-S mutants) (Fig. 4I). All variations of the MDFIC C terminus remained stably bound (fig. S16, A and B). An overlay of a snapshot from simulations shows that the acyl chains coincide with the additional MDFIC cryo-EM densities (fig. S16C). We compared all inter- actions that occurred in either unmodified or lipidated MDFIC monomers with those in the 5C-S mutant, which we know binds to but does not modulate PIEZO1 function (Fig. 4J and fig. S16). Interactions between the unmodified and 5C-S MDFIC were indistinguishable, which sup- ports the critical role of MDFIC lipidation in modulating PIEZO1 function. By contrast, the lipidated MDFIC had multiple distinct inter- actions with inner-helix residues (Fig. 4, J and K, and fig. S16, D and E). Some of these resi- dues within the inner helix are critical for in- activation, including Leu2475 (30), making this the likely pathway for functional modifi- cation. Given that MDFIC would need to stay bound throughout the PIEZO1 conformational cycle, we also examined whether MDFIC could bind to the flattened state of PIEZO1 (31). When MDFIC was aligned to the same pocket, sim- ulations showed that it interacted with the inner helix and remained stably bound (fig. S16, F to H). Discussion Using AC-MS, we have identified a family of Piezo-channel binding partners—the MyoD- family inhibitor proteins, MDFI and MDFIC— that fit all the criteria for Piezo-channel aux- iliary subunits. Although our study does not resolve full-length MDFIC or its native mem- brane topology, it does reveal that MDFIC binds to the PIEZO1 pore module through its con- served C terminus, which regulates channel inactivation. The regulation of Piezo-channel inactivation critically involves palmitoylation of the distal C-terminal of MDFIC. Because palmitoylation is a reversible lipid addition (29), this adds the potential for dynamic spatiotem- poral regulation of Piezo inactivation by MDFIC. Although we identified few other Piezo-channel binding-partner candidates with our “fibroblast- centric” screen, we speculate that the binding region of MDFIC could form a conserved bind- ing site for alternate membrane-associated Piezo regulators or auxiliary subunits. In support of this hypothesis, the PIEZO1 lysine residues that form salt bridges with MDFIC have been proposed to interact with other proteins (32). Despite PIEZO1 channels being able to func- tion as independent mechanosensors in sim- plified systems (2), ample evidence suggests that Piezos, particularly PIEZO2, may receive force through molecular tethers (33). Given the location of binding, it seems unlikely that MDFIC could act as a tethering molecule. It remains to be seen how ubiquitous this regulatory mechanism is; MDFIC and MDFI are absent from many cell types with rapidly in- activating PIEZO1 channels, including LNCaP (9) and N2A (1), but are expressed in others, including fibroblasts (10, 12, 13) and endothe- lial cells that exhibit slower PIEZO1 inactiva- tion (8, 11, 14). We isolated cardiac fibroblasts from mice that harbored a truncated MDFIC lacking the C terminus (19), in which PIEZO1 exhibited faster inactivation than in wild-type Zhou et al., Science 381, 799–804 (2023) 18 August 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E fibroblasts. Thus, MDFIC- or MDFI-mediated regulation of Piezo channels could be widely used. Despite our study not defining a PIEZO1- MDFIC complex in lymphatic endothelial cells, the similar lymphatic phenotypes associated with loss of function of MDFIC (19) and PIEZO1 (25, 26) means that this interaction may unearth molecular aspects underlying lymphatic vas- cular disease. Furthermore, both MDFIC and MDFI bind to transcription factors (19, 34) [in- cluding GATA2, a master regulator of lymphatic valve development (35)]. Whether PIEZO1/2 channels influence this aspect of their func- tion is unknown but may lay the foundation for the unveiling of a direct mechanosignaling pathway by means of Piezos to transcription through GATA2 or other transcription factors. Thus, our structural and functional data not only reveal a family of Piezo-channel auxiliary subunits but may also reveal a conserved bind- ing site for other membrane-associated pro- teins that modulate Piezo-channel function and, in doing so, the basis for an alternate Piezo- dependent mechanosignaling pathway. RE FE RENCES AND N OT ES 1. B. Coste et al., Science 330, 55–60 (2010). 2. R. Syeda et al., Cell Rep. 17, 1739–1746 (2016). 3. C. D. Cox et al., Nat. Commun. 7, 10366 (2016). 4. M. Yao et al., Sci. Adv. 8, eabo1461 (2022). 5. G. Vaisey, P. Banerjee, A. J. North, C. A. Haselwandter, R. MacKinnon, eLife 11, e82621 (2022). J. R. Holt et al., eLife 10, e65415 (2021). 6. 7. S. Yang et al., Nat. Commun. 13, 7467 (2022). 8. S. Wang et al., Nat. Commun. 11, 2303 (2020). 9. R. Maroto, A. Kurosky, O. P. Hamill, Channels (Austin) 6, 290–307 (2012). 10. N. M. Blythe et al., J. Biol. Chem. 294, 17395–17408 (2019). 11. J. I. Del Mármol, K. K. Touhara, G. Croft, R. MacKinnon, eLife 7, e33149 (2018). 12. E. Glogowska et al., Cell Rep. 37, 110070 (2021). 13. D. Jakob et al., J. Mol. Cell. Cardiol. 158, 49–62 (2021). 14. J. Shi et al., Cell Rep. 33, 108225 (2020). 15. L. O. Romero et al., Nat. Commun. 10, 1200 (2019). 16. A. C. Dolphin, J. Physiol. 594, 5369–5390 (2016). 17. D. Yan, S. Tomita, J. Physiol. 590, 21–31 (2012). 18. S. Thébault, F. Gachon, I. Lemasson, C. Devaux, J. M. Mesnard, J. Biol. Chem. 275, 4848–4857 (2000). 19. A. B. Byrne et al., Sci. Transl. Med. 14, eabm4869 (2022). 20. M. Karlsson et al., Sci. Adv. 7, eabh2169 (2021). 21. R. Syeda et al., eLife 4, e07369 (2015). 22. C. M. Chen, N. Kraut, M. Groudine, H. Weintraub, Cell 86, 731–741 (1996). 23. M. Moroni, M. R. Servin-Vences, R. Fleischer, O. Sánchez-Carranza, G. R. Lewin, Nat. Commun. 9, 1096 (2018). 24. J. Wu et al., Cell Rep. 21, 2357–2366 (2017). 25. K. Nonomura et al., Proc. Natl. Acad. Sci. U.S.A. 115, 12817–12822 (2018). 26. E. Fotiou et al., Nat. Commun. 6, 8085 (2015). 27. E. O. Anderson, E. R. Schneider, J. D. Matson, E. O. Gracheva, S. N. Bagriantsev, Cell Rep. 23, 701–708 (2018). 28. Y. Xie et al., Sci. Rep. 6, 28249 (2016). 29. A. Main, W. Fuller, FEBS J. 289, 861–882 (2022). 30. W. Zheng, E. O. Gracheva, S. N. Bagriantsev, eLife 8, e44003 (2019). 31. X. Yang et al., Nature 604, 377–383 (2022). 32. J. Wang et al., Cell Rep. 38, 110342 (2022). 33. C. Verkest, S. G. Lechner, Curr. Opin. Physiol. 31, 100625 (2023). 34. L. Snider et al., Mol. Cell. Biol. 21, 1866–1873 (2001). 35. J. Kazenwadel et al., J. Clin. Invest. 125, 2979–2994 (2015). ACKN OWLED GMEN TS We thank B. Martinac, E. Perozo, J. Xu, and the Patapoutian Lab for useful scientific input and Y. Zhang and Y. Yuan for mass spec support. We thank the Cryo-EM Center at the Interdisciplinary Research Center on Biology and Chemistry of the Shanghai Institute of Organic Chemistry for help with data collection. Funding: This work was supported by Australian Research Council (ARC) Future Fellowship FT220100159 (C.D.C.); STI2030-Major Projects 2022ZD0207400, Shanghai Municipal of Science and Technology Project 20JC1419500, and Shanghai Municipal Science and Technology Major Project 2019SHZDZX02 (Y.Z.); National Health and Medical Research Council Ideas Grant 2021260 and the Lymphatic Malformation Institute (N.L.H.); ARC DP200100860 (B.C.); and Medical Research Future Fund grant GHFM76777 (H.S.S.). Research was undertaken with the assistance of resources and services from the National Computational Infrastructure, supported by the Australian Government. Author contributions: Z.Z., M.Y., and D.C. generated cell lines and performed biochemical and imaging experiments. Z.Z., N.B., and C.D.C. performed electrophysiology. G.A.S., D.L.S., H.S.S., V.J., R.P.H., and N.L.H. generated animal models and isolated cells. Z.Z. and J.V.L. generated reagents. X.M. and Y.Z. performed cryo-EM structural studies. B.C. and Y.L. performed molecular dynamics simulations. C.D.C. and Y.Z. conceived of and supervised the project. All authors wrote and approved the manuscript. Competing interests: C.D.C. and Z.Z. note an Australian provisional patent application #2023902096 entitled “Piezo channel regulators.” Data availability: The composite map of the PIEZO1-MDFIC complex has been deposited in the Electron Microscopy Data Bank under the accession code EMD-35577. The composite map may contain artificial features near the boundaries of the masks. The consensus map, masked refined cap map, masked refined TMD map, and the map of PIEZO1-MDFIC (C240) have been deposited under the accession codes EMD-36241, EMD-36242, EMD-36243, and EMD-36244. The atomic coordinates have been deposited in the Protein Data Bank under the accession code 8IMZ. All data are available in the main text or the supplementary materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh8190 Materials and Methods Supplementary Text Figs. S1 to S16 Tables S1 to S3 References (36–47) Submitted 16 March 2023; accepted 19 July 2023 10.1126/science.adh8190 Zhou et al., Science 381, 799–804 (2023) 18 August 2023 6 of 6
10.1126_science.adi1071
Post date 6 June 2024 Retraction On 8 December 2023, Science published the Research Article “Drought sensitivity in mesic forests heightens their vulnerability to climate change” (1). In January 2024, Stefan Klesse, William Anderegg, John Abatzoglou, and Margaret E. K. Evans alerted the authors to errors in an R script on which they had relied to calculate potential evapotranspiration and climatic water deficit using a heat load–modified Thornthwaite equation. In response, the authors updated to the newest, corrected version of this script and reran their analysis. Although many results remained unchanged after these corrections, the statistical significance of some conclu- sions, and the results of some robustness tests, did change. In addition, the paper’s primary result was not robust to the use of an alternate climate input dataset that calculated potential evapotranspiration using the Penman-Monteith equation. As a result, all authors agree to retract the paper. The authors thank S. Klesse, W. Anderegg, J. Abatzoglou, and M. E. K. Evans for quickly detecting and raising these issues. Robert Heilmayr1,2*, Joan Dudney1,2, Frances C. Moore3 1Environmental Studies Program, University of California, Santa Barbara, Santa Barbara, CA, USA. 2Bren School of Environmental Science and Management, University of California, Santa Barbara, Santa Barbara, CA, USA. 3Department of Environmental Science and Policy, University of California, Davis, Davis, CA, USA. *Corresponding author. Email: rheilmayr@ucsb.edu REFEREN C E S A N D N OTES 1. R. Heilmayr, J. Dudney, F. C. Moore, Science 382, 1171 (2023). SCIENCE science.org RETRACTION POST DATE • 6 JUNE 2024 1 10.1126/science.adq4537 RES EARCH FOREST ECOLOGY Drought sensitivity in mesic forests heightens their vulnerability to climate change Robert Heilmayr1,2*, Joan Dudney1,2, Frances C. Moore3 Climate change is shifting the structure and function of global forests, underscoring the critical need to predict which forests are most vulnerable to a hotter and drier future. We analyzed 6.6 million tree rings from 122 species to assess trees’ sensitivity to water and energy availability. We found that trees growing in wetter portions of their range exhibit the greatest drought sensitivity. To test how these patterns of drought sensitivity influence vulnerability to climate change, we predicted tree growth through 2100. Our results suggest that drought adaptations in arid regions will partially buffer trees against climate change. By contrast, trees growing in the wetter, hotter portions of their climatic range may experience unexpectedly large adverse impacts under climate change. F orests cover approximately 30% of Earth’s land surface (1), absorb more carbon than all other terrestrial ecosystems (2, 3), and provide trillions of dollars’ worth of eco- system services (4, 5). Climate change, however, is causing pronounced changes in forests, including drought-induced diebacks and declines in productivity. Scientists warn that forests will continue to shift from carbon sinks to sources as climate change–related dis- turbances increase (6–8). To effectively man- age and respond to these changes, there is a critical need for accurate predictions that de- tail which forests will be most vulnerable to drier and hotter conditions (9, 10). Precise estimates of forest vulnerability to drought depend on an understanding of how spatial variation in drought sensitivity inter- acts with anticipated changes in drought expo- sure (Fig. 1) (11–13). Will forests located near their dry-range edge be pushed beyond their physiological limits or can drought adapta- tion in drier regions insulate forests against hotter droughts? Answering this question rests on a clear understanding of trees’ relative sen- sitivity to dry conditions across their range. However, past research exploring variation in drought sensitivity has yielded contradic- tory results, supporting either a “dry-range– sensitive” or a “drought-naïve” hypothesis. The dry-range–sensitive hypothesis predicts that trees growing in drier portions of their range are the most drought sensitive. Trees growing in resource-poor conditions—for ex- ample, shallow soils, steep slopes, and aspects exposed to greater solar radiation—are often more climate and drought sensitive (14–16). Similarly, many studies have observed greater reductions in radial growth during drought at drier sites than at more-mesic sites (15, 17–22). One explanation for these patterns is that in- creasing aridity in drier regions could exceed species’ physiological limits, leading to marked declines in growth in response to drought (10, 23). By contrast, the drought-naïve hypothesis predicts that trees growing in wetter portions of their range are the most drought sensitive (10). Drought-naïve effects have been found in forests in the Pacific Northwest (24), pine forests in the Eastern Mediterranean (25), European beech trees (Fagus sylvatica) (26), and ponderosa pine (Pinus ponderosa) pop- ulations in North America (27). Higher drought sensitivity in wetter regions may reflect fewer drought adaptations as a result of lower drought exposure. Long-term drought exposure, which is often higher in arid regions, can shift pheno- typic plasticity (26, 28), tree density (29, 30), and heritable traits that together confer greater drought resilience (31). Reconciling whether dry-range–sensitive or drought-naïve dynam- ics are more prevalent at a global scale can advance our ability to predict forest vulner- ability to climate, which is fundamental to 1Environmental Studies Program, University of California, Santa Barbara, Santa Barbara, CA, USA. 2Bren School of Environmental Science and Management, University of California, Santa Barbara, Santa Barbara, CA, USA. 3Department of Environmental Science and Policy, University of California, Davis, Davis, CA, USA. *Corresponding author. Email: rheilmayr@ucsb.edu Sensitivity A ) D W C Δ / I W R Δ ( y t i v i t i s n e S 0.000 -0.025 -0.050 -0.075 -0.100 B D W C n i e g n a h C 2 1 0 Exposure Vulnerability C 0.0 I W R n i e g n a h C -0.1 -0.2 -2 -1 0 Standardized aridity (Deviation from species mean) 1 2 -2 -1 0 Standardized aridity (Deviation from species mean) 2 1 -2 -1 0 Standardized aridity (Deviation from species mean) 1 2 H0: Consistent H1: Dry-range H2: Drought-naïve Fig. 1. Hypothesized vulnerability to climate change. (A) Three hypothesized patterns of heterogeneity in drought sensitivity across an aridity gradient: (H0) Trees show a consistent drought response across their species’ climatic range (consistent sensitivity; green line); (H1) trees growing in more-arid portions of their range are more drought sensitive (dry-range sensitive; gray line); and (H2) trees growing in more- mesic portions of their range are more drought sensitive (drought-naïve; purple line). Sensitivity (y axis) is defined as a change in annual growth (measured using RWI) in response to interannual fluctuations in water availability (measured using CWD). Standardized aridity (x axis) is defined as the historic average CWD at a given location relative to the mean CWD experienced across a species’ range. (B) Increases in CWD are widespread but anticipated to be highest in more-arid regions. Bar heights illustrate average increases in CWD CMIP5 models by 2100. (C) The interaction between sensitivity (A) and exposure (B) drives predicted changes in RWI, which we used as a measure of vulnerability. Colored lines and corresponding bars represent hypothesized patterns of vulnerability (change in RWI) when different patterns of sensitivity (A) interact with anticipated increases in drought exposure (B). Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 1 of 7 RETRACTED 6 June 2024. See Retraction. RES EARCH | R E S E A R C H A R T I C L E Data preparation Step 1 Step 2 Estimation Step 3 Step 4 Step 5 Step 6 Prediction Fig. 2. Methodological approach, using data for P. ponderosa as an illustration. Step 1: We extracted tree-ring width measurements from ITRDB sites (pink dots, with two example sites highlighted in purple and green) and downloaded geographic species ranges for all species in the ITRDB (P. ponderosa distribution shown in gray). Then, we detrended raw ring width measurements into RWI. Step 2: We downloaded global climate data and calculated annual CWD and PET, which we standardized for each species using the historic climate experienced across the species’ geographic range. The figure shows the distribution of mean PET and CWD between 1901 and 1980 across the ponderosa pine’s full range (contours) and within ITRDB sites (dots). Step 3: We estimated sensitivity within each site. Step 4: We estimated heterogeneity in sensitivity across historic standardized CWD and PET using site-level sensitivity estimates. Although this figure illustrates this approach for the subset of PIPO sites, the primary model specification discussed in the rest of this paper conducts this analysis across all species. Step 5: We predicted sensitivity across the geographic range of all species. Step 6: We predicted change in RWI as a function of predicted sensitivities and CMIP5 predictions of changes in CWD and PET. developing effective management strategies (32–34). We address this long-standing debate by (i) identifying where trees exhibit the greatest sensitivity to changes in water and energy availability and (ii) predicting where forests will be most vulnerable to climate change. To quantify trees’ sensitivity to weather fluctua- tions, we drew upon two large databases of annual tree-ring measurements: the Interna- tional Tree Ring Data Bank (ITRDB) and the US Forest Service’s Forest Inventory and Analy- sis (FIA) tree-ring dataset (35). The ITRDB data we analyzed includes 6.6 million annual observations of tree growth from 3775 sites spanning 122 species (listed in table S1). In addition to its geographic and taxonomic scope, an important benefit of the ITRDB is that the distribution of observed climatic condi- tions provides a fairly representative sample of macroscale climatic variation experienced across tree species’ full geographic ranges (fig. S2). However, the ITRDB exhibits known biases toward trees growing in marginal con- ditions that exhibit relatively strong climate signals (36, 37), which may overstate trees’ sensitivity to weather (15). To ensure that our results were not an artifact of ITRDB biases, we replicated our analysis using FIA tree cores. Though FIA data are more geographically and taxonomically limited than those of the ITRDB, FIA tree cores are drawn from a systematic sampling design, thereby enabling a more representative understanding of ecological processes. We analyzed these data using six steps (Fig. 2) (38). First, we detrended all tree-ring width measurements using standard dendrochrono- logy approaches to generate annual observations of ring width index (RWI), a standardized mea- sure of tree growth. Second, we characterized each tree species’ climatic niche by extracting historic weather data across the entirety of each species’ geographic range. We characterized weather and climate using potential evapo- transpiration (PET) and climatic water deficit (CWD), two variables drawn from climatic water balance equations that emphasize the inter- acting roles that water and energy availa- bility play in constraining tree species’ ranges (39–41). Specifically, PET describes the atmo- spheric demand for water from vegetation, which is primarily a function of the amount of energy available to plants. By contrast, CWD quantifies shortages in water availability rela- tive to this atmospheric demand. To enable comparisons across species, we standardized CWD and PET values using the distribution of historic CWD and PET experienced across the species’ full range (fig. S1). Third, for each ITRDB site, we regressed annual, tree-level observa- tions of RWI (step 1) on fluctuations in species- standardized weather (step 2) to estimate the sensitivity of individual sites to interannual Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 2 of 7 RETRACTED 6 June 2024. See Retraction. RES EARCH | R E S E A R C H A R T I C L E A T E P c i r o t s H i i ) n a e m s e c e p s m o r f n o i t a v e D i ( 2 1 0 -1 -2 B D W C o t y t i v i t i s n e s . d e r P C D W C o t y t i v i t i s n e s . d e r P 0.0 -0.1 -0.2 0.1 0.0 -0.1 Marginal effect of CWD 0.0 -0.1 -0.2 -0.3 Marginal effect of PET 0.2 0.0 -0.2 -0.4 D T E P c i r o t s H i i ) n a e m s e c e p s m o r f n o i t a v e D i ( 2 1 0 -1 -2 -2 -2 -2 -1 0 Historic CWD (Deviation from species mean) 1 -1 0 Historic CWD (Deviation from species mean) 1 -1 0 Historic PET (Deviation from species mean) 1 2 -2 E T E P o t y t i v i t i s n e s . d e r P 0.0 -0.2 -0.4 -0.6 2 -2 F T E P o t y t i v i t i s n e s . d e r P 0.2 0.0 -0.2 -0.4 2 -2 -1 0 Historic CWD (Deviation from species mean) 1 -1 0 Historic CWD (Deviation from species mean) 1 -1 0 Historic PET (Deviation from species mean) 1 2 2 2 Fig. 3. Variation in sensitivity to CWD and PET across species’ climatic niches. (A and D) Marginal effects of CWD (A) and PET (D) from site-specific, first-stage models, averaged within bins distributed across the standardized, historic CWD-PET climate space. First-stage coefficients represent the sensitivity of RWI to changes in CWD (high CWD sensitivity in dark red and low CWD sensitivity in gray). Sensitivity is measured as the change in RWI that would result from a 1-SD increase in CWD or PET. (B, C, E, and F) Second-stage models were used to summarize heterogeneity in sensitivity to CWD across a gradient of historic CWD (B) and PET (C) and heterogeneity in sensitivity to PET across historic CWD (E) and PET (F). Shading represents the 2.5 to 97.5% interquantile range from the block-bootstrapped second-stage model. changes in CWD and PET. Fourth, we quan- tified how the sensitivities estimated in step 3 varied across sites along a gradient of his- toric, species-standardized climate. Fifth, using this estimate of heterogeneity in sensitivity along a climatic gradient, we predicted cli- matic sensitivities across each species’ full geographic range. Finally, we estimated how climate change will shift tree growth across species’ ranges by combining predicted sen- sitivities (step 5) with climate predictions from the Representative Concentration Path- way 6.0 scenario (RCP 6.0) of the Coupled Model Intercomparison Project Phase 5 (CMIP5). To ensure that our final results reflected the combined uncertainty of steps 3 to 6 of this analysis, we propagated uncertainty using a combination of Monte Carlo simulation and block bootstrapping. Effects of CWD and PET on tree growth Most ITRDB sites exhibited positive growth re- sponses to increases in either water or energy availability (fig. S3), a pattern that is consis- tent with ecological theory and previous studies (9, 10, 24, 40). Specifically, a site that exper- ienced a 1 SD decline in CWD (relative to the distribution of CWD across the species’ full range) experienced a 3.2% decline (95% confi- dence interval: −10.1 to −1.1) in RWI in the same year. The site-level, marginal effect of CWD was significantly negative (p < 0.05) for 54.6% of sites, whereas only 10.6% of sites exhibited a significantly positive relationship (fig. S3A). By contrast, the relationship between PET and growth varied greatly across sites, with signif- icant positive relationships in 39.7% of sites and significant negative relationships in 25.3% of sites (fig. S3B). Site-level, distributed lag models provided further evidence that nega- tive CWD shocks and positive PET shocks led to increases in growth, with the strongest ef- fects occurring in the year of the shock rather than in lagged years (fig. S3, C and D). Drier sites exhibit less sensitivity to drought Consistent with the drought-naïve hypothesis, trees located in drier parts of their species’ climatic range exhibited less sensitivity to drought than those in more mesic sites (Fig. 3). For example, our estimates of site-level sensi- tivity to CWD indicate that drier-than-average sites experienced a 12.8% decline (−22.3 to −6.0%) in annual growth in response to a 1-SD increase in CWD (Fig. 3A). By contrast, wetter- than-average sites experienced a larger 22.1% decline (−34.4 to −11.2%) in growth from a 1-SD increase in CWD. To summarize this variability in CWD sensitivity, we modeled sensitivity as a quadratic function of historic CWD and PET (Fig. 3, B and C). The regression results highlight that observed sensitivity de- clined with increasing CWD [slope at the spe- cies’ mean, historic CWD = 0.015 (0.006 to 0.059)]. This slope indicates that a 1-SD in- crease in CWD leads to an approximately 1.5 percentage point greater decline in growth in sites located at their species’ mean CWD than in a site that is 1 SD drier. Trees also exhibited heterogeneity in their responses to annual increases in energy avail- ability (Fig. 3, D to F). In the most energy- limited portions of species’ ranges, trees often experienced large increases in growth in response Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 3 of 7 RETRACTED 6 June 2024. See Retraction. RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Predicted changes in CWD and PET. (A) Spatial distribution of the mean predicted change in CWD through 2100. (B and C) Predicted increases in CWD (B) and PET (C) are largest in regions of high historic CWD and PET. (D) Predicted changes in CWD would push most species’ historic ranges above their historic mean CWD. All plots compare changes between the historic (1970–2000) and end-of- century (2091–2100) predic- tions according to the average of CMIP5’s 47 model runs for the RCP 6.0 scenario. A B 0 0 1 2 − 1 9 0 2 n i D W C n a e M i ) n a e m s e c e p s m o r f n o i t i a v e D ( 20 15 10 5 0 −5 D 1.00 e g n a r f o n o i t r o p o r P 0.75 0.50 0.25 0.00 −5 C 0 0 1 2 − 1 9 0 2 n i T E P n a e M i ) n a e m s e c e p s m o r f n o i t i a v e D ( 15 10 5 0 −5 20 −5 0 5 10 15 Mean PET in 1970−2000 (Deviation from species mean) Change in CWD (mm/year) 1500 1000 500 0 Count of grid cells 5000 4000 3000 2000 1000 0 Change in CWD Beyond prior max 1−2 s.d. above prior mean 0−1 s.d. above prior mean Below prior mean 0 5 10 15 Mean CWD in 1970−2000 (Deviation from species mean) s ulu p o P a e Pic s u xin a r F rix a L a g u s t o d u e s P a g u s T s bie A s u Pin s u g a F ria a c u a r A s u r e nip u J ris a p y c e a m a h C uja h T s u r d e c o Lib s u g fa o h t o N s u c r e u Q s u r d e c alo C m diu o x a T n o r d n e d o r e r e c A ula t e B s u p r a c alo H s u d cla yllo h P xis a t o r h t A s u r d e c o r t s u A a e n a t s a C n o r d n e d ria e m o t p y r C Lirio s u r d e C a y o r z Fit a y r a C Pilg Genera to increased PET. For example, in the coldest 10% of sites (<1.5 SD below the species’ his- toric mean PET), growth increased 18.7% (8.7 to 25.1%) in response to a 1-SD increase in PET. However, this positive growth response to in- creased annual PET quickly declined across a gradient of historic PET (Fig. 3F). In sites with above-average historic PET, increased an- nual PET was associated with an insignificant decline in growth. Our main conclusion—that tree sensitivity to CWD decreases across an aridity gradient— is robust to a variety of alternate models and datasets (38). For example, fig. S6 contrasts estimates of heterogeneity derived from system- atically sampled, FIA tree-ring data against results derived from comparable ITRDB data. The similarity in results indicates that the drought-naïve effect we observed is unlikely to be the result of biases in ITRDB site selection. In addition, we found that the drought-naïve effect persists under a wide variety of alternate approaches to data processing and analysis (fig. S5). Finally, although drought-sensitivity patterns vary across the 10 most common genera in the ITRDB, most exhibited a significant drought-naïve pattern of CWD sensitivity, whereas none of the genera exhibited a sig- nificantly higher sensitivity in the dry portions of their range (fig. S4). The results provide Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 4 of 7 RETRACTED 6 June 2024. See Retraction. RES EARCH | R E S E A R C H A R T I C L E C Predicted change in CWD (std) E Predicted change in RWI Predicted sensitivity to CWD −0.05 −0.06 −0.07 −0.08 Predicted sensitivity to PET 0.15 0.10 0.05 0.00 A 2 ) n a e m 1 0 −1 −2 B 2 1 0 −1 −2 T E P c i r o t s H i i s e c e p s m o r f n o i t a v e D i ( ) n a e m i s e c e p s T E P c i r o t s H i m o r f n o i t i a v e D ( 3 2 1 D Predicted change in PET (std) 2.0 1.5 1.0 0.1 0.0 −0.1 −0.2 2 1 0 −1 −2 T E P c i r o t s H i ) n a e m i s e c e p s m o r f n o i t a v e D i ( 2 −2 −2 −1 1 Historic CWD 0 2 −2 −1 1 0 Historic CWD (Deviation from species mean) (Deviation from species mean) −1 0 Historic CWD (Deviation from species mean) 1 2 Fig. 5. Predicted changes in tree growth through 2100. (A and B) Predicted mean sensitivity of RWI to interannual changes in CWD (A) and PET (B) aggregated across the historic climatic ranges of all ITRDB species. Sensitivity is measured as the change in RWI that would result from a 1-SD increase in CWD or PET. (C and D) Forecasted changes in CWD (C) and PET (D) between CMIP5 RCP 6.0’s historic (1970–2000) and end-of-century (2090–2100) periods, measured in standard deviations. (E) Resulting predictions of the impacts of climate change on tree growth (RWI) across species’ historic climatic ranges. further support for our conclusion that CWD sensitivity declines along a gradient of his- toric CWD. Climatic changes across species ranges Tree species will generally face much warmer and drier conditions by 2100 (Fig. 4). Specif- ically, CMIP5 model ensembles predict that the entirety of every species’ range analyzed will experience increases in both mean PET and CWD (Fig. 4, B and C). On average, the mean PET across a species’ range is antici- pated to increase by 1.39 SD (0.45 to 3.05), whereas CWD is anticipated to increase by 1.41 SD (0.34 to 3.78). Supporting concerns about the emergence of “novel climates” (42, 43), 11% of the average species’ range will be drier than anywhere in that species’ historic, cli- matic range (Fig. 4D). For some species (e.g., Pinus pinea and Quercus faginea), more than half of their current range is projected to be drier than the driest parts of their historic range. Although the greatest increases in tem- perature are expected to occur in northern latitudes as a result of snow and ice albedo feedbacks (44), CMIP5 models predict higher rates of drying in warmer, lower latitudes (Fig. 4A). An important mechanism driving this pattern is the nonlinear effect of temperature on the water-holding capacity of the atmo- sphere (45, 46). As a result of this nonlinear effect, PET is anticipated to increase more rapidly in hotter locations (Fig. 4C). Fore- casted changes in annual precipitation across most ITRDB sites are not sufficient to offset warming-induced increases in atmospheric water demand, leading to pervasive increases in CWD. These increases in CWD are also antic- ipated to be most pronounced in dry and hot locations. The largest growth declines are predicted to occur in wet and hot regions Combining our estimates of sensitivity to CWD and PET with predicted changes in climate, we estimate that by 2100, tree growth will decline by 10.4% (−37.3 to +0.7%). More forests are predicted to experience declines in growth than increases in growth. Specifically, 51.2% of grid cells (representing the combination of a species and location) are predicted to experi- ence a significant (>95% of Monte Carlo itera- tions) decline in growth. By contrast, only 0.6% of grid cells are predicted to experience a significant increase in growth. These results highlight that a hotter, drier planet will likely lead to major shifts in the health of global forests and their ability to sequester carbon. Predictions of RWI changes through 2100 highlight that climate change may have a sur- prisingly pronounced and negative impact on trees growing in wetter, warmer portions of their range, whereas cooler parts of species’ dry- range edges may be unexpectedly resilient to climate change. Figure 5 illustrates how spatial variation in sensitivity (Fig. 5, A and B) inter- acts with climate change exposure (Fig. 5, C and D) to yield heterogeneity in predicted RWI declines (Fig. 5E). The wetter but hotter-than- average portions of species’ ranges are pro- jected to experience a 17.2% decline in growth (−51.9 to −1.4%). By contrast, drier but cooler- than-average locations are predicted to experi- ence a smaller 11.0% decline in growth (−33.9% to +0.3%). Grid cells predicted to experience in- creases in growth are primarily located in very energy-limited parts of species’ historic ranges. Capturing heterogeneity in sensitivity in eco- logical models is important to accurately pre- dict the impacts of climate change. For example, a neutral model, in which predicted sensitivity to PET and CWD is constant across species’ climatic ranges, would yield contrasting results to our model, which allows for spatial hetero- geneity. Specifically, the neutral model would likely underestimate the negative impact of cli- mate change in wetter regions but overestimate the negative impact in drier regions when com- pared with our model (fig. S7). These results highlight that the spatial heterogeneity in trees’ climatic sensitivities will likely play a critical role in mediating the response of the world’s forests to climate change. Discussion Trees growing in wet and warm portions of their range will likely experience the greatest declines in growth under climate change. This Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 5 of 7 RETRACTED 6 June 2024. See Retraction. RES EARCH | R E S E A R C H A R T I C L E pattern of vulnerability captures the inter- action of heterogeneity in both exposure and sensitivity—historically wet forests are more sensitive to drier conditions, whereas hot forests will be exposed to the greatest in- creases in CWD. Though drought-induced de- clines in tree growth frequently occur in dry regions (15, 19, 47), our results support the hypothesis that the vulnerability of wetter forests to climate change is greatly under- appreciated (48). The drought-naïve pattern exhibited across ITRDB and FIA sites is likely the result of long-term adaptations to local climatic condi- tions. Drought adapted traits—for example, higher resistance to cavitation, lower leaf-area index, higher water-use efficiency, and wider hydraulic safety margins (49, 50)—often e- merge in regions that experience higher aridity and drought frequency (51–53). In theory, drought adaptations could occur at the scale of individuals (e.g., phenotypic plasticity) or local communities (e.g., tree density) or through evolutionary processes that occur across gen- erations. Recent research, however, suggests that most climate adaptation develops over longer timescales than an individual tree’s life span (54). This suggests that genetic adap- tation is an important mechanism that pro- tects trees from climate change, but it also highlights the risk that rapid warming may outpace the rate of adaptation. Although sensitivity patterns for many gen- era are consistent with the drought-naïve effect (fig. S4), the sign, magnitude, and signif- icance of this effect varies across taxa. Some of this variability is likely due to differences in tree traits (e.g., turgor loss point, xylem struc- ture, leaf size) and mechanisms of acclimation (e.g., canopy thinning, trait plasticity, shifts in water allocation) (23, 48). Interestingly, Quercus and Taxodium, the two genera that exhibited opposite (although insignificant) patterns of drought sensitivity, both have hotter species ranges than the other commonly sampled gen- era (fig. S4). This suggests that taxonomic differences in drought sensitivity may be partly explained by trade-offs between investments in heat and drought tolerance (55), as well as the physiological limits of climate adaptation. Future research that disentangles the primary mechanisms of drought and heat tolerance will improve estimates of species-level vulner- ability to climate change. The prevalence of drought-naïve effects is relevant for understanding climate change impacts across many scales of biological or- ganization. Organisms and ecosystems with prior exposure to extreme weather (e.g., droughts and heat waves) are often less sensitive to extreme events than entities that lack past exposure. This pattern has been found across diverse systems, including soil microbial com- munities (56), insect species (57), bird popula- tions (58), agricultural production (59), human mortality (60), and even biomes (61). Though the mechanisms that enable adaptation vary by system, relatively consistent patterns of sen- sitivity underscore the important role that adaptation could play in moderating the ef- fects of climate extremes. Because we quantified vulnerability using tree growth rather than mortality, our analysis provides incomplete insights into patterns of mortality under climate change. Declines in RWI have been shown to predict tree mortal- ity (62, 63), and the primary mechanisms of drought-induced mortality—hydraulic failure, carbon starvation, and defensive failure (23)— can be linked to changes in tree growth (64). Therefore, our predictions of growth declines in mesic regions suggest that increases in tree mortality may be surprisingly widespread. Nevertheless, declines in RWI can be an un- reliable predictor of mortality (64, 65). A tree’s age, genotype, functional traits, and mecha- nisms of acclimation (64, 66, 67) are all im- portant determinants of growth responses to drought, which can decouple the relationship between RWI and mortality. Importantly, our predictions of vulnerability might understate the risk of mortality at dry-range edges if there are thresholds in the relationship between tree growth and mortality (68). For example, trees likely require a minimum amount of growth to survive (68); therefore, the same percent- age growth decline could be more harmful for slow-growing trees located at their dry-range edge than for trees growing in mesic areas (10). Future research that characterizes these nonlinear responses to drought could improve predictions of climate impacts (42, 43). Finally, our findings have important policy implications, both for conserving forests and managing terrestrial carbon sinks. Policy-makers who seek to protect forests from climate change may need to expand the focus of conservation interventions beyond species’ dry-range edges. By contrast, drought adaptations in popula- tions from drier regions could be useful for management interventions, including assisted migration into wetter regions. Additionally, although dynamic global vegetation models are increasingly sophisticated (69–72), very few, if any, account for the spatial variation in the climate sensitivity of carbon capture. Fail- ing to account for drought-naïve effects can lead to overestimates of the resilience of car- bon sinks in wetter regions. Improving fore- casts of climate-change impacts on terrestrial carbon sinks by accounting for the variability in drought sensitivity will facilitate more-effective global policies aimed at reducing carbon emis- sions (73, 74). RE FERENCES AND NOTES 2. W. R. L. Anderegg et al., Science 368, eaaz7005 (2020). 3. Y. Pan et al., Science 333, 988–993 (2011). 4. R. Costanza et al., Nature 387, 253–260 (1997). 5. R. Costanza et al., Glob. Environ. Change 26, 152–158 (2014). 6. W. A. Kurz et al., Nature 452, 987–990 (2008). 7. O. L. Phillips et al., Science 323, 1344–1347 (2009). 8. K. Giles-Hansen, X. Wei, Environ. Res. Lett. 17, 044049 (2022). 9. C. D. Allen et al., For. Ecol. Manage. 259, 660–684 (2010). 10. W. R. L. Anderegg, L. D. L. Anderegg, K. L. Kerr, A. T. Trugman, Glob. Change Biol. 25, 3793–3802 (2019). 11. M. Lindner et al., For. Ecol. Manage. 259, 698–709 (2010). 12. 13. T. J. Brodribb, J. Powers, H. Cochard, B. Choat, Science 368, I. Aubin et al., Ecosphere 9, e02108 (2018). 261–266 (2020). 14. X. Zhang, P. Lv, C. Xu, X. Huang, T. Rademacher, Agric. For. Meteorol. 308–309, 108611 (2021). 15. S. Klesse et al., Nat. Commun. 9, 5336 (2018). 16. H. C. Fritts, D. G. Smith, J. W. Cardis, C. A. Budelsky, Ecology 46, 393–401 (1965). 17. C. Desplanque, C. Rolland, F. H. Schweingruber, Trees 13, 218–227 (1999). 18. F. Lebourgeois, C. B. K. Rathgeber, E. Ulrich, J. Veg. Sci. 21, 364–376 (2010). 19. A. P. Williams et al., Proc. Natl. Acad. Sci. U.S.A. 107, 21289–21294 (2010). 20. H. Liu et al., Glob. Change Biol. 19, 2500–2510 (2013). 21. T. Kolář, P. Čermák, M. Trnka, T. Žid, M. Rybníček, Agric. For. Meteorol. 239, 24–33 (2017). 22. L. D’Orangeville et al., Glob. Change Biol. 24, 2339–2351 (2018). 23. N. G. McDowell et al., Nat. Rev. Earth Environ. 3, 294–308 (2022). 24. D. Mildrexler, Z. Yang, W. B. Cohen, D. M. Bell, Remote Sens. Environ. 173, 314–325 (2016). 25. D. Helman, I. M. Lensky, D. Yakir, Y. Osem, Glob. Change Biol. 23, 2801–2817 (2017). 26. L. Cavin, A. S. Jump, Glob. Change Biol. 23, 362–379 (2017). 27. I. M. McCullough, F. W. Davis, A. P. Williams, For. Ecol. Manage. 402, 223–233 (2017). 28. B. Choat et al., Nature 491, 752–755 (2012). 29. K. E. Gleason et al., Ecosphere 8, e01849 (2017). 30. A. Aguirre, M. del Río, S. Condés, For. Ecol. Manage. 411, 90–100 (2018). 31. R. López et al., Ann. Bot. 111, 1167–1179 (2013). 32. T. M. Brooks et al., Science 313, 58–61 (2006). 33. M. I. Williams, K. Dumroese, in Forest Conservation in the Anthropocene: Science, Policy, and Practice, V. A. Sample, R. P. Bixler, C. Miller, Eds. (Univ. Press of Colorado, 2016), pp. 113–123. 34. W. J. Wang, H. S. He, F. R. ThompsonIII, J. S. Fraser, W. D. Dijak, Landsc. Ecol. 32, 1399–1413 (2017). 35. R. J. DeRose, J. D. Shaw, J. N. Long, J. For. 115, 283–291 (2017). 36. F. Babst et al., Quat. Sci. Rev. 197, 1–20 (2018). 37. S. Zhao, N. Pederson, L. D’Orangeville, J. Biogeogr. 46, 355–368 (2019). 38. Materials and methods are available as supplementary materials. 39. N. L. Stephenson, Am. Nat. 135, 649–670 (1990). 40. N. Stephenson, J. Biogeogr. 25, 855–870 (1998). 41. D. D. Ackerly et al., Front. Ecol. Environ. 18, 288–297 (2020). 42. J. W. Williams, S. T. Jackson, Front. Ecol. Environ. 5, 475–482 (2007). 43. C. R. Mahony, A. J. Cannon, T. Wang, S. N. Aitken, Glob. Change Biol. 23, 3934–3955 (2017). 44. Q. You et al., Earth Sci. Rev. 217, 103625 (2021). 45. D. D. Breshears et al., Front. Plant Sci. 4, 266 (2013). 46. A. Park Williams et al., Nat. Clim. Change 3, 292–297 (2013). 47. J. Su et al., For. Ecol. Manage. 482, 118841 (2021). 48. C. D. Allen, D. D. Breshears, N. G. McDowell, Ecosphere 6, 1–55 (2015). 49. C. Pritzkow, V. Williamson, C. Szota, R. Trouvé, S. K. Arndt, Tree Physiol. 40, 215–229 (2020). 50. J. M. R. Peters et al., Glob. Change Biol. 27, 3620–3641 (2021). 1. Food and Agriculture Organization of the United Nations, Global Forest Resources Assessment 2005; https://www.fao. org/3/a0400e/a0400e00.htm [accessed 26 August 2020]. 51. Y. Osem, K. O’Hara, Forestry 89, 338–349 (2016). 52. T. Chauvin, H. Cochard, V. Segura, P. Rozenberg, For. Ecol. Manage. 444, 9–20 (2019). Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 6 of 7 RETRACTED 6 June 2024. See Retraction. RES EARCH | R E S E A R C H A R T I C L E 53. R. López, F. J. Cano, N. K. Martin-StPaul, H. Cochard, B. Choat, New Phytol. 230, 497–509 (2021). 54. J. Gantois, Glob. Change Biol. 28, 6002–6020 (2022). 55. R. M. Marchin et al., Glob. Change Biol. 28, 1133–1146 (2022). 56. R. Ochoa-Hueso et al., Glob. Change Biol. 24, 2818–2827 (2018). 57. C. García-Robledo, E. K. Kuprewicz, C. L. Staines, T. L. Erwin, W. J. Kress, Proc. Natl. Acad. Sci. U.S.A. 113, 680–685 (2016). 58. B. L. Bateman et al., Landsc. Ecol. 30, 1095–1110 (2015). 59. E. E. Butler, P. Huybers, Nat. Clim. Chang. 3, 68–72 (2013). 60. T. Carleton et al., Q. J. Econ. 137, 2037–2105 (2022). 61. S. M. Vicente-Serrano et al., Proc. Natl. Acad. Sci. U.S.A. 110, 52–57 (2013). 62. A. J. Das, J. J. Battles, N. L. Stephenson, P. J. van Mantgem, Can. J. For. Res. 37, 580–597 (2007). 63. J. J. Camarero, A. Gazol, G. Sangüesa-Barreda, J. Oliva, S. M. Vicente-Serrano, J. Ecol. 103, 44–57 (2015). 64. M. Cailleret et al., Glob. Change Biol. 23, 1675–1690 (2017). 65. N. Bircher, M. Cailleret, H. Bugmann, Ecol. Appl. 25, 1303–1318 (2015). 66. A. T. Trugman, New Phytol. 234, 21–27 (2022). 67. N. L. Stephenson, A. J. Das, N. J. Ampersee, B. M. Bulaon, J. L. Yee, J. Ecol. 107, 2383–2401 (2019). 68. J. Au et al., Glob. Change Biol. 29, 5652–5665 (2023). 69. G. Krinner et al., Global Biogeochem. Cycles 19, 2003GB002199 (2005). 70. B. Smith et al., Biogeosciences 11, 2027–2054 (2014). 71. M. Fernández-Martínez et al., Nat. Clim. Chang. 9, 73–79 (2019). 72. M. Forkel et al., Sci. Rep. 9, 18757 (2019). 73. A. Favero, A. Daigneault, B. Sohngen, Sci. Adv. 6, eaay6792 (2020). 74. L. Nascimento et al., Clim. Policy 22, 158–174 (2022). 75. R. Heilmayr, Replication code and data for: Drought sensitivity in mesic forests heightens their vulnerability to climate change. Open Science Framework (2023); https://doi.org/10.17605/ OSF.IO/RHJCF. 76. H. D. Grissino-Mayer, H. C. Fritts, Holocene 7, 235–238 (1997). 77. Climatic Research Unit (University of East Anglia) and Met Office, CRU TS Version 4.04 (2020); https://crudata.uea.ac. uk/cru/data/hrg/cru_ts_4.04/ [accessed 9 March 2023]. 78. World Climate Research Programme, Coupled Model Intercomparison Project Phase 5; https://esgf-node.llnl.gov/ search/cmip5/. ACKN OWLED GMEN TS We thank R. Huezo for help parsing the ITRDB database and M. Leung for support in compiling our database of species range maps. We express our appreciation to the many contributors to the ITRDB. In addition, we thank S. Klesse and J. Shaw for providing access to the US Forest Service’s National Forest Inventory and Analysis Program’s Interior West tree-ring dataset. Funding: J.D. acknowledges support from the David H. Smith Conservation Research Fellowship. F.C.M. acknowledges support from the National Science Foundation (NSF award no. 1924378). Author contributions: R.H., J.D., and F.C.M. all contributed to the conceptualization, data curation, formal analysis, writing (original draft), and writing (review and editing) of this paper. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All code used to clean data, conduct the analysis, and generate figures and results is available at this paper’s Open Science Framework (OSF) repository (75). This repository also includes the preprocessed data necessary to reproduce all results, figures, and tables. Raw input data are available from the ITRDB (76), the Climate Research Unit (77), and the Coupled Model Intercomparison Project (78). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi1071 Materials and Methods Figs. S1 to S7 Table S1 References (79–91) MDAR Reproducibility Checklist Submitted 7 April 2023; accepted 26 October 2023 10.1126/science.adi1071 Heilmayr et al., Science 382, 1171–1177 (2023) 8 December 2023 7 of 7 RETRACTED 6 June 2024. See Retraction.
10.1126_science.adi1563
RES EARCH 3D PRINTING Self-enhancing sono-inks enable deep-penetration acoustic volumetric printing Xiao Kuang1†, Qiangzhou Rong2†, Saud Belal2†, Tri Vu2, Alice M. López López1, Nanchao Wang2, Mehmet Onur Arıcan1, Carlos Ezio Garciamendez-Mijares1, Maomao Chen2, Junjie Yao2*, Yu Shrike Zhang1* Volumetric printing, an emerging additive manufacturing technique, builds objects with enhanced printing speed and surface quality by forgoing the stepwise ink-renewal step. Existing volumetric printing techniques almost exclusively rely on light energy to trigger photopolymerization in transparent inks, limiting material choices and build sizes. We report a self-enhancing sonicated ink (or sono-ink) design and corresponding focused-ultrasound writing technique for deep-penetration acoustic volumetric printing (DAVP). We used experiments and acoustic modeling to study the frequency and scanning rate–dependent acoustic printing behaviors. DAVP achieves the key features of low acoustic streaming, rapid sonothermal polymerization, and large printing depth, enabling the printing of volumetric hydrogels and nanocomposites with various shapes regardless of their optical properties. DAVP also allows printing at centimeter depths through biological tissues, paving the way toward minimally invasive medicine. T hree-dimensional (3D) printing is at- tracting increasing attention with its ability to directly fabricate geometrically complex constructs for prototypes (1), high-performance materials (2, 3), multi- material parts (4, 5), flexible electronics (6, 7), medical devices (8), and engineered tissues (9, 10). Various printing modalities, such as extrusion printing, inkjet printing, stereolithog- raphy, and powder-bed fusion, have been de- veloped to print different materials, including thermoplastics, liquid photoinks, and solid polymer powders, among others (11, 12). These printing methods use light or photothermal heating as energy sources to trigger selective material solidification in a layer-by-layer fash- ion. A build platform controlled by a linear translation stage is usually required to support the stepwise material solidification. Emerg- ing printing techniques accompanied by new photocurable inks have been developed to improve printing speed (13, 14), printing res- olution (15–17), and printout functionality (18, 19). Volumetric printing that creates 3D constructs without a build platform or an ink-renewal step can substantially improve printing speed and surface quality (20–23). The existing volumetric printing techniques use light to achieve selective photopolymer- ization in the volumes of optically transparent inks (24–26). However, the light attenuation by inks themselves, the presence of functional additives (e.g., photoabsorbers and fillers), or/and already-cured parts have imposed con- 1Division of Engineering in Medicine, Department of Medicine, Brigham and Women’s Hospital, Harvard Medical School, Cambridge, MA 02139, USA. 2Department of Biomedical Engineering, Duke University, Durham, NC 27708, USA. *Corresponding author. Email: junjie.yao@duke.edu (J.Y.); yszhang@bwh.harvard.edu (Y.S.Z.) †These authors contributed equally to this work. straints on the material choices and the build sizes for light-based volumetric printing. Al- though infrared (IR) light can be used to im- prove light penetration to several millimeters (26–28), it remains technically challenging to deliver light deep into optically scattering media, such as biological tissues. Therefore, light-based volumetric printing has intrinsic limitations for application in deep-penetration digital manufacturing schemes and, further, in minimally invasive fabrication scenarios (29, 30). Compared with light, ultrasound waves (<10 MHz) can penetrate >100 times deeper into optically scattering materials and thus hold promise for depositing energy to trigger polymerization at depths (31). By means of an ultrasound bath or horn-based reactors, ultra- sound waves can generate reactive oxygen spe- cies (ROS; i.e., hydroxyl and peroxide radicals) by cavitation of water, enabling vinyl mono- mer polymerization for hydrogel formation in several to tens of minutes (32, 33). Ultrasound waves can also be focused into a small volume using a focused ultrasound (FUS) transducer. A FUS transducer can generate acoustic waves with positive and negative pressures alternat- ing at megahertz frequency and propagating along the depth direction, and the high acous- tic energy can be delivered into the focal zone with high precision. Previously, cavitation- based ultrasound printing was achieved by curing a polydimethylsiloxane resin (34). How- ever, a build platform was required, and only relatively simple geometries could be printed, because the intense acoustic streaming by the high acoustic pressure disturbed the local ink at the focus region. Here, we report phase-transition viscoelastic sonicated inks (hereafter, sono-inks) that simul- taneously allow deep acoustic penetration, low acoustic streaming, and rapid sonothermally induced radical polymerization, collectively enabling deep-penetration acoustic volumet- ric printing (DAVP) (Fig. 1A). DAVP takes ad- vantage of rapid material solidification by the sonothermal effect of the FUS focus in a visco- elastic sono-ink, which provides the building voxel to construct 3D objects without the need for a build platform. In DAVP, the FUS waves deliver deep-penetration acoustic energies with pressures up to several tens of mega- pascals to the local region at a distance of up to 64 mm (focal length) (Fig. 1B). The small oval-shaped FUS focal zone (full width at half maximum of the acoustic pressure field: 0.3 to 0.7 mm) is further narrowed by the nonlinear acoustic propagation effect at high acoustic pressure (35), collectively facilitating fast, high-resolution printing (Fig. 1C and fig. S1). Consequently, DAVP allows us to print geometrically complex materials precisely and volumetrically, even through nontransparent and optically scattering materials. DAVP principle and self-enhancing sono-ink design Traditionally, ultrasound-mediated cross-linking of the vinyl-based hydrogel precursors is slow, because of the low concentration of ROS gen- erated by ultrasound-induced cavitation (see supplementary text in the supplementary mate- rials). Besides, ROS can be rapidly quenched or diluted by robust acoustic streaming. Our simulation results showed that high-viscosity fluids could significantly reduce the acoustic streaming velocity (Fig. 1D and fig. S2). How- ever, highly viscous ink feedstocks usually exhibit high acoustic attenuations and thus substantially reduce acoustic penetration (36). We hypothesized that a meticulously designed multicomponent viscoelastic sono-ink should suppress acoustic streaming while facilitating a fast sonothermal effect and thus trigger rapid and spatial radical polymerization of vinyl precursors for a deep-penetration fabri- cation scheme. To formulate such a proof-of-concept sono- ink, we selected a vinyl oligomer of poly (ethylene glycol) diacrylate (PEGDA) as the base component, agar microparticles as the rheology modifier, poly(N-isopropyl acrylam- ide) (PNIPAm) as the self-enhancing acoustic absorber, and ammonium persulfate (APS) as the thermal initiator (fig. S3). Compared with PEGDA ink, PEGDA-based sono-ink (PEGDA/ agar/PNIPAm at, for example, 20/10/3 wt/wt/ wt % ratio) showed enhanced viscosities ow- ing to the coil-to-globule phase transition of PNIPAm [transition temperature (Tt) = 34° to 36°C] (Fig. 1E and fig. S4). The viscosity of the sono-ink increased 92-fold, from 2.0 to 185.3 Pa·s, when heated from 25°C to 40°C at low shearing (0.05 Hz) (fig. S4C). Mean- while, the sono-ink showed prominent shear Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Working principle of DAVP and design of self-enhancing sono-ink. (A) Scheme showing acoustic printing of constructs by selective curing of sono-ink using deep-penetration FUS. The sonothermal effect enhanced by the phase-transitioning acoustic absorber triggers the decomposition of the initiator for local polymerization of acrylate oligomers into polymer networks at the heating zone. (B) Simulated acoustic pressure field in water for the 3.41-MHz FUS transducer with an acoustic power of 50 W, using the nonlinear acoustic model. (C) Lateral pressure distribution of the 3.41-MHz FUS transducer with an output voltage (Vp) of 192 V. The shaded regions show the full width at half maximum of the positive and negative pressures (0.3 to 0.7 mm). (D) Simulated maximum velocity in the focal region versus the dynamic viscosity of the sono-ink. Inset acoustic streaming simulation shows the velocity field of fluids with a viscosity of 2.5 Pa·s. (E) Oscillation temperature sweep shows the enhanced moduli and viscosity of PEGDA-based sono-ink at a transition temperature (33° to 36°C) due to coil-to-globule transition of PNIPAm. (F) Acoustic streaming in different fluids at the focal region of 3.41-MHz FUS for 2-s exposure: 20 wt % PEGDA (i) and PEGDA/agar/PNIPAm sono-ink (ii). Scale bars: 5 mm. (G) Acoustic properties of the PEGDA/agar/PNIPAm sono-ink and its key components under 3.41-MHz FUS at 25° and 37°C: acoustic attenuation coefficient in linear scale (i) and penetration depth in log scale (ii). (H) Peak temperature in sono-ink at the heating zone near the FUS focus as a function of exposure time, at an environment temperature of 24°C. (I) Gelation time measured by rheology as a function of the curing temperature for PEGDA-based sono-ink supplemented with 1.0 w/w% APS. Orange squares represent the experiment data, and the solid blue line is the fitting curve by the Arrhenius law. thinning, as reflected by an 87% reduction in viscosity under 100-Hz shearing at 25°C (fig. S4). This sono-ink design concept was gen- eralizable to different formulations, includ- ing those using vinyl oligomers from natural polymers [e.g., gelatin methacryloyl (GelMA)] (fig. S4) or different phase-transition poly- mers with tunable transition temperatures (Tt,offset = 20.9° to 38.5°C) (fig. S5 and table S1), as well as formulations adding various nanoparticles (table S2). Our sono-ink de- sign resolved the long-standing dilemma be- tween acoustic penetration depth and acoustic streaming. On the one hand, the shear thin- ning facilitated deep acoustic penetration under high-frequency acoustic waves. On the other hand, the viscosity enhancement by phase tran- sition substantially reduced acoustic stream- ing. In contrast to the violent streaming in the PEGDA solution, the self-enhancing sono-ink exhibited negligible fluid flow at the FUS focus (Fig. 1F, fig. S6, and movie S1), as supported by the simulation (Fig. 1D). To investigate the sono-ink’s self-enhanced acoustic attenuation effect, we measured the acoustic attenuation coefficient (a) of the proof- of-concept sono-ink and key components at various temperatures (fig. S7). Because of the phase transition, the a of the PNIPAm (3 wt %) aqueous solution and PEGDA-based ink at the ultrasound frequency of 3.41 MHz increased with the temperature, for example, by 600 and 100%, respectively, from 25°C to 37°C (Fig. 1G). The a of the sono-ink (containing both PEGDA and PNIPAm) was 25.2 nepers per meter (Np m−1) at a temperature of 37°C, which was 86-fold larger than that of 20 wt % PEGDA alone but still much lower than the a of soft biological tissues under the same conditions (table S3). Despite the self-enhanced acoustic attenuation, the acoustic waves could still achieve a large penetration depth (Dp = 1/a) of 40 mm into the sono-ink at 37°C (Fig. 1G). As a direct result of self-enhanced acoustic attenuation and viscosity, the phase-transition sono-ink exhibited a fast and self-enhanced heating effect. The PNIPAm solution and the sono-ink could be rapidly heated up to between Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E 60° and 80°C at the heating zone after sec- onds of FUS exposure at mild input powers (<100 W) (fig. S8 and movie S2). PNIPAm played a critical role in achieving fast sono- heating, in contrast to the negligible heating of the 20 wt % PEGDA solution. To further in- vestigate the self-enhanced heating effect, the heating-zone temperatures as a function of FUS exposure time were monitored by an IR thermal camera during FUS exposure at dif- ferent environmental temperatures (Tenv). At a Tenv of 24°C—far below the Tt, offset of PNIPAm —the peak temperature at the heating zone first showed a slow heating rate (4.8°C s−1) and then, after the temperature leap at 36°C, a much-enhanced heating rate (11.3°C s−1), as a result of the phase transition (Fig. 1H). The slow heating period, called the induction period, was reduced from 0.5 s to 0.1 s by raising the Tenv to 33°C (close to Tt, offset) and entirely elim- inated at Tenv = 35°C (>Tt, offset) (fig. S9 and movie S3). In our printing technique, Tenv was set slightly lower than Tt, offset to allow for confined sonoheating at the FUS focus. It is worth noting that, as in thermal-based tissue ablation (37), our self-enhancing sono-ink en- abled a prominent sonothermal effect under a high-duty cycle of 90%. We did not observe FUS-induced cavitation activities, as validated by active cavitation mapping (fig. S10 and movie S4). Therefore, it was concluded that FUS-induced cavitation is not responsible for the sono-ink solidification. The self-enhanced sonothermal effect was leveraged to trigger the radical polymeriza- tion of vinyl oligomers for fast gelation. The gel-point conversion of the PEGDA-based sono- ink was ~0.3, as measured by Fourier trans- form infrared spectroscopy (fig. S11). In the presence of APS, the sono-inks proceeded with (sono)thermal-induced gelation, as illu- strated by the rheological measurement (fig. S12). The temperature-sensitive gelation time (tgel) followed the Arrhenius law, revealing a high activation energy (171 kJ mol−1) (eq. S10). Consequently, long shelf life (tgel = 167 days at 4°C) and on-demand fast curing (tgel = 1.9 s at 80°C) were simultaneously achieved (Fig. 1I). DAVP printing-resolution characterizations We developed a 3D FUS printer, which was equipped with a FUS transducer at three ultra- sound frequencies (2.05, 3.41, and 6.86 MHz), a 3D motorized translation stage, and a print- ing control system (fig. S13; see materials and methods in the supplementary materials for further details). We first studied the single- point printing resolution of DAVP. Upon FUS exposure in the sono-ink, a whitening region formed and quickly expanded beyond the FUS heating zone (fig. S14 and movie S5), as PNIPAm simultaneously acted as the self-enhancing acoustic absorber and the temperature indi- cator (38). IR thermal images of the sono-ink surface showed a clear temperature increase forming a heating zone (Fig. 2A). The heated surface area (diameter of up to 4 mm) was larger than the focal zone of the FUS transducer (0.45 mm at 3.41 MHz) because of the rapid thermal diffusion. Upon cooling, a piece of opaque solid formed at the center of the heat- ing zone, and the uncured sono-ink returned to its original color. By comparing the curing size with the heating-zone temperature pro- file, we found that the curing temperature threshold was Tc = 67°C for the PEGDA-based sono-ink supplemented with 0.5 w/w% APS. Because of heat diffusion by the low-power cumulative FUS, the curing size increased with input power (or peak acoustic pressure) and exposure time (Fig. 2B and fig. S15). The non- linear acoustic effect resulted in elevated acous- tic intensity at the FUS focus by increasing the peak pressure and reducing the focal region (39, 40), as shown by our nonlinear acoustic modeling (fig. S16). The simulation also cap- tured the nonlinear acoustic propagation in the sono-ink, exhibiting weak reflection and scattering by different boundaries (such as plastics and tissues) in the far field (fig. S17), and negligible acoustic reflection and scatter- ing at the interface formed by the sono-ink phase transition (fig. S18). Using the numer- ically modeled temperature map and the measured Tc, we investigated the increasing single-point curing size (0 to 4 mm in diam- eter) as a function of peak acoustic pressure (35 to 55 MPa) and exposure time (0 to 5 s) (Fig. 2C, fig. S19, and movie S6). Tc of the sono-inks could be readily reduced from 67°C to 62°C by increasing the APS concentration, because of the enhanced reaction rate (fig. S20 and eq. S8). Additionally, increasing the PEGDA con- centration (40 wt %) increased the apparent Tc of the sono-ink, likely owing to additional heating by the exothermic effect of polymeri- zation and reduced viscosity (fig. S21). We further investigated the printing resolu- tions of DAVP by continuously scanning a line under different printing settings. An expand- ing heated region instantly formed at the front of the scanning FUS focus, as a result of the thermal diffusion within the ink, which was also confirmed by sonothermal modeling (Fig. 2D, fig. S22, table S4, and movie S7). The sono- ink was cross-linked only at the center of the heating zone, where the temperature increased above the curing threshold, forming aniso- tropic filaments. We also investigated the de- pendence of the curing size on several key parameters, including the scanning speed of the FUS focus, the FUS frequency and power, and thermal diffusion (figs. S23 to S25 and movie S8). Modeling results showed that the thin ink tank (6 mm) caused heat accumula- tion at the tank boundary, whereas a thicker tank (>10 mm) allowed for normal thermal diffusion around the focal region (fig. S26). Therefore, we used the 10-mm-thick ink tank to quantify the curing sizes (or printing reso- lutions) (Fig. 2E). First, increasing the scan- ning speed can improve the printing resolution. We observed overcuring in the longitudinal dimension (larger than design) at a slow scan- ning speed with all three ultrasound frequen- cies (2.05, 3.41, and 6.86 MHz), owing to excess heat accumulation. In comparison, undercur- ing (lower than design) was observed at high scanning speeds because of inadequate tem- perature increase at the two ends. The in- plane curing size was determined by the joint effect of the FUS scanning speed and thermal diffusion. For instance, the in-plane curing size at 3.41 MHz was improved from 7.6 mm to 1.6 mm by increasing the scanning speed from 0.4 mm s−1 to 0.8 mm s−1. Similarly, the axial curing size was mostly determined by the FUS depth of focus and thermal diffusion. As an example, the 6.86-MHz FUS had a smaller depth of focus (1.69 mm) than the 2.05-MHz FUS (5.58 mm), but the sono-ink had a 10-fold acoustic absorption at 6.86 MHz and thus much higher heating efficiency. Eventually, the axial curing size ranged from 5.0 to 9.9 mm for the investigated printing parameters. The depth of focus here is defined as the size of the focal zone along the acoustic axis, in which the acoustic pressure drops to half of the peak value (table S5). The printing resolution of DAVP can be flexibly adjusted by optimizing the FUS scan- ning speed, frequency, and power. Generally, the printing anisotropy (ratio of the axial and in-plane curing size) increased with the print- ing speed, mostly as a result of the decreased in-plane curing size (Fig. 2F). For example, when printing at 0.8 mm s−1 by the 6.86-MHz FUS, the cross section of printed filaments ex- hibited an oval shape with an axial size of 8.8 mm, which was 100% larger than the in- plane curing size (Fig. 2, G and H). The sono- thermal curing mechanism also led to different microstructures within the FUS heating zone. The cured hydrogels displayed sub-micrometer pores in the PEGDA/PNIPAm matrix due to polymerization-induced phase separation, as shown by scanning electron microscopy (SEM) (Fig. 2I). SEM images suggested melting of agar microparticles owing to a high tempera- ture of >85°C at the center of the heating zone (Fig. 2I, panel ii). However, the agar micro- particles maintained a granular shape outside the center of the focal zone at various printing speeds and frequencies, similar to that achieved by bulk heating at 60° to 70°C (figs. S28 and S29). These results further excluded the pres- ence of FUS-induced cavitation and confirmed the heat accumulation–based curing mecha- nism. Meanwhile, the enhanced surface qua- lity of the printed filaments was confirmed by the uniform and smooth surface (fig. S29D). Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Characterizations of DAVP printing resolution. (A) Typical surface temperature profile near the FUS focal region. The IR thermal image shows the in-plane temperature map (i) and temperature diagram at the cross section (ii). The inset shows the cured solid after 3-s FUS exposure (3.41 MHz) with the curing threshold temperature of 67°C. (B) In-plane curing size of printed solids versus FUS exposure time using PEGDA-based sono-ink (0.5 w/w% APS) at different Vp, as labeled. (C) Color-contour of modeled in-plane curing size at 3.41 MHz as a function of acoustic pressure and exposure time. (D) Snapshots of photographs (i) and IR thermal images (ii) near the FUS focal zone, as well as modeled temperature map on the focal plane (iii) at different time points during line printing using PEGDA-based sono-ink (1.0 w/w% APS) (FUS frequency: 3.41 MHz; printing speed: 1.5 mm s−1). (E) Curing sizes for single-line DAVP printing using different printing speeds at various FUS frequencies as labeled: longitudinal size (i), in-plane size (ii), and z-axial size (iii). (F) Axial-to-lateral curing size ratio as a function of printing speed using various FUS frequencies, as labeled. (G) Photographs of printed filaments by single-line DAVP (6.86-MHz FUS at a speed of 0.8 mm s−1): in-plane size (i) and axial size (ii). (H) Cross-sectional micrographs of printed filaments in (G) with three different marked regions. (I) SEM images of dried hydrogels in (H) at different regions: top (i), center (ii), and bottom (iii). Scale bars: 2 mm [(A), (D), (G), (H)] and 2 mm (I). When performing multipath printing, the axial printing size (or printing thickness) in- creased, as a result of repeated FUS exposure along the acoustic axis at the scanned regions; however, the printing lengths were less sensi- tive to the number of printing paths (fig. S30). Additionally, the printing performance of DAVP was independent of the line-scanning direc- tion of the 3D FUS printer. There were no significant differences in printing fidelity and mechanical properties (e.g., Young’s moduli: 1.4 ± 0.1, 1.2 ± 0.3, and 1.3 ± 0.1 MPa) using different scanning directions (0°, 45°, and 90°, respectively) (fig. S31). Moreover, the printed sample at the 0° infill angle showed signifi- cantly higher stiffness than that by bulk heat- ing (0.9 ± 0.1 MPa), likely because of the high temperature–induced fast matrix curing and agar welding. DAVP of volumetric constructs and material generality Deep penetration is the primary advantage of DAVP over conventional light-based 3D print- ing (Fig. 3A). For example, in a black-dyed sono-ink (0.5 w/w%), the penetration depths for the 2.05-, 3.41-, and 6.86-MHz ultrasound waves were 295.2, 86.8, and 28.2 mm at room temperature, respectively, which were, respec- tively, 600, 180, and 60 times as large as those for the 405-nm light (Dp = 0.48 mm) (Fig. 3B). As such, a large curing depth of 24 mm in the dyed sono-ink was achieved in 26 s at an axial scanning speed of 1 mm s−1 and 3.41-MHz FUS, whereas only a thin solid (2.4 mm) in the dye- stained photoink was obtained after photo- curing for 165 s (fig. S32). The large penetration depth of DAVP, regardless of the optical prop- erties of the ink, allowed us to volumetrically Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. DAVP performance and material generality. (A) Schematic illustrations for photocuring of photoinks (i) and FUS-curing of sono-inks (ii) with different penetration capabilities. (B) Penetration depth as a function of black-dye content for UV light (405 nm) and FUS at different frequencies, as labeled. (C to F) Design models (i), top-view (ii), and tilt-view photographs (iii) of printed constructs of various shapes using PEGDA-based sono-ink: honeycomb (C), vessel network (D), 3D hand (E), and spider (F). (G to J) Photographs of printed constructs using various sono-inks: multicolor three-part gear set (i) and a three-part wheel set (ii) by dye-stained PEGDA-based sono-inks consisting of different colors (G), vascular network model by a fluorescence-stained PEGDA-based sono-ink (0.1 w/w% rhodamine B) (H), tree-shaped model by a PEGDA-based nanocomposite sono-ink (10 wt % nanoclay) (I), and layered heart-shaped model by a GelMA-based protein sono-ink (J). (K to M) In vitro cytocompatibility of GelMA-based sono-inks and cured hydrogels: representative fluorescence micrographs of live (green)/dead (red) cells (NIH/3T3 fibroblasts) after 30-min exposure to DPBS (control) and GelMA- based sono-ink (K), quantitative cellular viability values of NIH/3T3 fibroblasts after 30-min exposure to GelMA-based sono-ink (i) and at day 7 after seeding on GelMA- based hydrogel (ii) (L), fluorescence micrographs showing F-actin staining of cells at day 7 after seeding on the GelMA/PNIPAm hydrogel using two cell types: human mesenchymal stem cells (hMSCs) (i) and human umbilical vein endothelial cells (HUVECs) (ii). F-actin, red; nucleus, blue. ns, no significant difference. Scale bars: 10 mm [(C) to (J)] and 100 mm [(K) and (M)]. print nontransparent 2D and 3D hydrogel con- structs with different sizes and geometrical complexities, including letters and lattices with sharp corners and spirals and vessels with smooth surfaces and transitions (fig. S33, table S6, and movie S9). For example, a 10-layer honeycomb and a vascular network were printed at 3.41 MHz (Fig. 3, C and D, and table S6). A complex 3D hand model (67 mm by 53 mm by 10 mm) and a spider model (52 mm by 43 mm by 10 mm) were also printed at 6.86 MHz (Fig. 3, E and F; fig. S34; and table S6). Using dye-stained PEGDA-based sono-inks, we printed opaque-colored hydrogel compos- ites for prototypes, including a set of three gears of different sizes and an assembled wheel set of three colored parts (Fig. 3G and fig. S35). Altering the phase-transition polymers with higher Tt, such as poly(N-isopropylmethacrylamide) (PNIPMAm), could facilitate printing at body temperature (37°C) (fig. S35D). Additionally, fluorescence-stained sono-ink (0.1 wt % rho- damine B) was used to print a high-fidelity branched vascular network shape (81 mm by Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. DAVP for proof-of-concept through-tissue printing and minimally invasive therapy. (A) Scheme of minimally invasive therapy by through-tissue manufacturing of scaffolds on target lesions and tissues. (B to D) Tissue phantoms with marked thicknesses (i) and printed objects with associated digital models (ii) for through-tissue manufacturing: a bone-shaped model printed under porcine belly (B), a lattice printed under porcine liver (C), and a heart-shaped model printed under porcine kidney (D). Printing was performed with 2.05-MHz FUS at 0.6 mm s−1. (E and F) DAVP for LAA closure for a 12-mm-thick goat heart: scheme and printing process at different time points (E), and stretching of closed LAA (F). (G and H) DAVP for defective bone reconstruction: scheme and photographs showing the procedures of manufacturing a nanocomposite filler material on a bone defect through 10-mm-thick skin and muscle (G), and ultrasound images of the bone defect (i) and the reconstructed bone (ii) (H). (I to K) DAVP for chemotherapy drug delivery for a 14-mm-thick porcine liver: scheme and photographs showing the procedure of printing drug-eluting hydrogels using a doxorubicin-loaded sono-ink on a liver lesion (I), cross-sectional photograph showing the printed hydrogel-liver interface (J), and fluorescence micrograph showing drug diffusion from the printed drug-eluting filler hydrogel to the liver at 1500-s contact (K). Scale bars: 10 mm [(B) to (J)] and 200 mm (K). 70 mm by 3 mm) (Fig. 3H). Further, a nano- composite sono-ink consisting of 10 wt % nanoclay was exploited to print a four-layer tree shape (2 mm in thickness) (Fig. 3I). DAVP was also used to print protein-based biomate- rials, as suggested by the tough-hydrogel heart model using the GelMA-based sono-ink (Fig. 3J and fig. S35, E and F). Both PEGDA- and GelMA-based sono-inks and their key compo- nents exhibited zero to low cytotoxicity using NIH/3T3 fibroblasts as a model cell line. High cell viability (>99%) similar to the control [using Dulbecco’s phosphate-buffered saline (DPBS)] was observed after direct and indirect expo- sure to the sono-ink for up to 30 min (Fig. 3, K and L, and fig. S36). Moreover, similar to the control cultured in a petri dish, the GelMA- based hydrogel enabled high viability (>99%) and healthy attachment and proliferation of postseeded mammalian cells of different types (Fig. 3M and fig. S37), indicating its favorable bioactivity. DAVP for through-tissue printing and constructive minimally invasive medicine As a proof of concept, we applied DAVP for high-speed and high-resolution through-tissue manufacturing and minimally invasive medi- cine (Fig. 4A). First, we illustrated the ex vivo through-tissue printing using soft tissues of different types and dimensions (fig. S38 and table S7). For through-tissue printing, thick tissues (up to 17 mm thick) were placed on top of the sono-ink chamber in the FUS near field. We printed a bone-shaped construct at 2.05 MHz through an ex vivo porcine tissue phantom consisting of a skin layer (3 mm), a fat layer (5 mm), and a muscle layer (7 mm) (Fig. 4B). High-fidelity honeycombs were printed through a 15-mm-thick porcine tissue con- sisting of skin and muscle or a 17-mm-thick porcine liver tissue (Fig. 4C). Similarly, a hol- low heart-shaped model was printed through a 17-mm-thick porcine kidney tissue (Fig. 4D and movie S10). Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E Nonvalvular atrial fibrillation is a prevalent cardiovascular disease related to the left atrial appendage (LAA) (41, 42). Open-chest surgery or transcatheter procedures can seal off the LAA to reduce the risk of thromboembolism. However, surgical LAA closure is severely in- vasive, and the treatment is often incomplete (41). We demonstrated proof-of-concept DAVP- assisted LAA closure (Fig. 4E). We delivered the sono-ink through a catheter to the LAA of an ex vivo goat heart placed in the printing cham- ber. The sono-ink was then solidified using the 3D FUS printer at 3.41 MHz through a 12-mm- thick heart wall. Precise FUS focus scanning en- abled selective curing of the sono-ink within the entire LAA volume while sparing surrounding heart tissues (fig. S39 and movie S11). After treatment, the cured hydrogels completely oc- cluded the LAA and bonded well with the tissue wall, which could tolerate reasonable distortions that mimicked the heart beating (Fig. 4F). We further explored the potential of the DAVP technique for tissue reconstruction and regeneration, such as treating large bone de- fects (43). As an illustration, the PEGDA/agar/ PNIPMAm nanocomposite sono-ink consist- ing of 5 w/w% hydroxyapatite (HAp) nanopar- ticles was formulated to print a bone scaffold for a hypothetical bone loss treatment (Fig. 4G and movie S12). We used a chicken leg to create a fibula bone defect model (1 cm long). After injecting the nanocomposite sono-ink, we printed a beam-shaped HAp-laden com- posite material through the skin and muscle tissues (10 mm thick) to reshape the defec- tive volume. The nanocomposite material could reconstruct the bone with seamless bonding to the native parts without influencing the surrounding tissues, as confirmed by ultra- sound imaging (Fig. 4H). We also demonstrated DAVP for therapeu- tic drug delivery by printing drug-eluting filler hydrogels on liver lesions (Fig. 4I). Doxorubi- cin, a clinical chemotherapy drug for treating a wide range of cancers, such as breast cancer and hepatocellular carcinoma, was used as a model drug (44). PEGDA/agar/PNIPMAm sono- ink consisting of 1 mg ml−1 of doxorubicin was formulated for printing chemotherapy drug- eluting hydrogels at the liver lesion sites at 37°C using FUS at various frequencies (fig. S40). DAVP enabled through-tissue printing and selective curing, as shown by the intimate hydrogel-tissue interfacial bonding (front side) and negligible burning of the intervening tis- sue (back side) (Fig. 4J and movie S13). The drug in the hydrogel gradually released and diffused into the liver tissue (effective diffu- sivity: 8.7 × 108 cm2 s−1), forming an ~3-mm- thick effective therapeutic layer [5.8 mg ml−1 concentration (44)] within 7 days by modeling (Fig. 4K and fig. S41). We anticipated that this method could be used as postablative chemo- therapy to improve cancer treatment. We note that relatively high FUS energies were needed for through-tissue printing, main- ly because of the acoustic attenuation by the intervening tissues. The acoustic energy loss might possibly result in overheating of the in- tervening tissues. To mitigate the risk of tissue overheating and to improve future in vivo printing efficiency, we developed a confocal- DAVP system in which the 3D FUS printer used two FUS transducers aligned in a cross- beam pattern, with their acoustic axes 90° apart and their foci overlapping (fig. S42). Taking advantage of the energy superposition of two acoustic foci, confocal-DAVP can achieve the curing temperature threshold in the com- bined foci, using roughly half of the output energy from each FUS transducer. Such a con- focal configuration has two advantages: (i) the acoustic energy deposition in the local inter- vening position is reduced by ~50%, and thus the overheating risk is minimized; (ii) the printing resolution and speed can be improved, owing to better sonothermal confinement within the overlapped foci. Indeed, the axial printing resolution was improved to 0.7 mm, and the printing speed reached 8 mm s−1 (fig. S42). Accordingly, complex models, including a leaf-shaped structure (55 mm by 35 mm) and a vessel-like network (75 mm by 25 mm), were successfully printed using the confocal-DAVP, taking only 20 and 83 s, respectively (fig. S43, table S8, and movie S14). Conclusions Leveraging the deep-penetration capability of FUS waves, low acoustic streaming, and rapid sonopolymerization of the viscoelastic self- enhancing sono-inks, we have developed a DAVP technique that can volumetrically build constructs with high printing fidelity and resolution in the absence of a build platform. The use of a thermally responsive adaptive acoustic absorber resolves the conflict be- tween acoustic streaming and deep penetra- tion upon FUS exposure. The self-enhancing sono-ink and nonlinear acoustic propaga- tion collectively enhanced the sonothermal heating at the FUS focus for fast and se- lective material solidification as the building voxel. The heat accumulation–based curing mechanism resulted in anisotropic printing resolution at a millimeter scale, which may be further improved by optimizing printing parameters of FUS frequency and scanning speed and by using the confocal dual-transducer configuration. The deep penetration of FUS waves allows the volumetric fabrication of opaque (nano)composites and printing through centimeter-thick tissues that are not attain- able through state-of-the-art light-based print- ing techniques (table S9). The self-enhancing sono-ink design can be generalized for differ- ent systems, greatly expanding the materials library for acoustic printing techniques. REFERENCES AND NOTES J. R. Tumbleston et al., Science 347, 1349–1352 (2015). 1. 2. S. Gantenbein et al., Nature 561, 226–230 (2018). 3. X. Zheng et al., Science 344, 1373–1377 (2014). 4. M. A. Skylar-Scott, J. Mueller, C. W. Visser, J. A. Lewis, Nature 575, 330–335 (2019). 5. K. Liu, R. Sun, C. Daraio, Science 377, 975–981 (2022). 6. H. Wei et al., Nat. Commun. 12, 2082 (2021). 7. H. Yuk et al., Nat. Commun. 11, 1604 (2020). 8. Y. Y. C. Choong et al., Nat. Rev. Mater. 5, 637–639 (2020). 9. L. Moroni et al., Nat. Rev. Mater. 3, 21–37 (2018). 10. A. Lee et al., Science 365, 482–487 (2019). 11. S. C. Ligon, R. Liska, J. Stampfl, M. Gurr, R. Mülhaupt, Chem. Rev. 117, 10212–10290 (2017). 12. M. L. Bedell, A. M. Navara, Y. Du, S. Zhang, A. G. Mikos, Chem. Rev. 120, 10744–10792 (2020). 13. D. A. Walker, J. L. Hedrick, C. A. Mirkin, Science 366, 360–364 (2019). 14. M. P. de Beer et al., Sci. Adv. 5, eaau8723 (2019). 15. S.-F. Liu et al., Science 377, 1112–1116 (2022). 16. F. Han et al., Science 378, 1325–1331 (2022). 17. N. M. Larson et al., Nature 613, 682–688 (2023). 18. G. Villar, A. D. Graham, H. Bayley, Science 340, 48–52 (2013). 19. X. Kuang et al., Sci. Adv. 5, eaav5790 (2019). 20. M. Shusteff et al., Sci. Adv. 3, eaao5496 (2017). 21. P. N. Bernal et al., Adv. Mater. 31, e1904209 (2019). 22. B. E. Kelly et al., Science 363, 1075–1079 (2019). 23. M. Xie et al., Nat. Commun. 14, 210 (2023). 24. M. Regehly et al., Nature 588, 620–624 (2020). 25. D. Loterie, P. Delrot, C. Moser, Nat. Commun. 11, 852 (2020). 26. S. N. Sanders et al., Nature 604, 474–478 (2022). 27. J. Wong et al., Adv. Mater. 35, e2207673 (2023). 28. J. Zhu, Q. Zhang, T. Yang, Y. Liu, R. Liu, Nat. Commun. 11, 3462 (2020). 29. M. Lee, R. Rizzo, F. Surman, M. Zenobi-Wong, Chem. Rev. 120, 10950–11027 (2020). 30. Y. Chen et al., Sci. Adv. 6, eaba7406 (2020). 31. T. G. McKenzie, F. Karimi, M. Ashokkumar, G. G. Qiao, Chemistry 25, 5372–5388 (2019). 32. B. Rokita, J. M. Rosiak, P. Ulanski, Macromolecules 42, 3269–3274 (2009). 33. Z. Gao et al., ACS Macro Lett. 8, 1285–1290 (2019). 34. M. Habibi, S. Foroughi, V. Karamzadeh, M. Packirisamy, Nat. Commun. 13, 1800 (2022). 35. R. T. Beyer, Phys. Acoust. 2, 231–264 (1965). 36. L. Claes et al., Measurement 184, 109919 (2021). 37. P. V. Yuldashev et al., IEEE Trans. Ultrason. Ferroelectr. Freq. Control 68, 2837–2852 (2021). 38. A. Rahimzadeh, M. Rutsch, M. Kupnik, R. V. Klitzing, Langmuir 37, 5854–5863 (2021). 39. A. Bhargava, V. C. Meesala, M. R. Hajj, S. Shahab, Appl. Phys. Lett. 117, 064101 (2020). 40. S. Haddadi, M. T. Ahmadian, Sci. Iran. 25, 2087–2097 (2018). 41. D. R. Holmes Jr., T. J. Simard, A. M. Killu, M. A. Alkhouli, Mayo Clin. Proc. 97, 1525–1533 (2022). 42. H. Ueno, T. Imamura, S. Tanaka, N. Fukuda, K. Kinugawa, J. Cardiol. 81, 420–428 (2023). 43. A. Stahl, Y. P. Yang, Tissue Eng. Part B Rev. 27, 539–547 (2021). 44. L. Barraud et al., J. Hepatol. 42, 736–743 (2005). 45. X. Kuang et al., Self-enhancing sono-inks enable deep- penetrating acoustic volumetric printing, Zenodo (2023); https://doi.org/10.5281/zenodo.8422050. AC KNOWLED GME NTS This work was performed in part at the Center for Nanoscale Systems (CNS) at Harvard University, a member of the National Nanotechnology Infrastructure Network (NNIN), which is supported by the National Science Foundation under award number ECS0335765. We also thank D. A. Weitz from Harvard University for rheometer access, and P. Zhong from the Department of Mechanical Engineering and Materials Science at Duke University for assistance with acoustic cavitation mapping. Funding: This work is supported by the National Institutes of Health (R21EB025270, R01HL153857, R01HL165176, R01HL166522, and R01CA282451), the National Science Foundation (CBET-EBMS- 1936105), and the Brigham Research Institute, all to Y.S.Z.; National Institutes of Health (NIH) [R21EB027981, R21EB027304, RF1NS115581 (BRAIN Initiative), R01NS111039, R01EB028143, R21EB027981, and R01EB031629], National Science Foundation Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E CAREER award 2144788, and Chan Zuckerberg Initiative Grant 2020-226178, all to J.Y. M.O.A. acknowledges funding support from the TÜBİTAK 2214-A Program (1059B141801395). Author contributions: X.K., J.Y., and Y.S.Z. conceived of the idea. X.K. designed the materials and methods. Q.R., S.B., X.K., M.C., T.V., and C.E.G.-M. designed the printer setup. X.K., A.M.L.L., and M.O.A. prepared the materials and conducted material characterizations. Q.R., S.B., T.V., N.W., and X.K. performed the printing and acoustic characterizations. Q.R. and X.K. developed the numerical models and conducted the simulations. X.K., Q.R., S.B., T.V., N.W., and C.E.G.-M. analyzed the results and processed visualizations. X.K., J.Y., and Y.S.Z. wrote the manuscript, with input from all authors. Y.S.Z. and J.Y. supervised the study. Competing interests: Y.S.Z. consults for Allevi by 3D Systems and sits on the scientific advisory board and holds options of Xellar, neither of which, however, participated in or biased this work. Data and materials availability: All data needed to evaluate the conclusions in the study are present in the paper and/or the supplementary materials. G-codes for printing and simulation codes for diffusion analysis are provided on Zenodo (45). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses- journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi1563 Materials and Methods Supplementary Text Figs. S1 to S43 Tables S1 to S9 References (46–81) MDAR Reproducibility Checklist Movies S1 to S14 Submitted 6 April 2023; accepted 13 October 2023 10.1126/science.adi1563 Kuang et al., Science 382, 1148–1155 (2023) 8 December 2023 8 of 8
10.1126_science.ade4245
RES EARCH QUANTUM SIMULATION Direct observation of nonlocal fermion pairing in an attractive Fermi-Hubbard gas Thomas Hartke*, Botond Oreg, Carter Turnbaugh, Ningyuan Jia, Martin Zwierlein The Hubbard model of attractively interacting fermions provides a paradigmatic setting for fermion pairing. It features a crossover between Bose-Einstein condensation of tightly bound pairs and Bardeen-Cooper-Schrieffer superfluidity of long-range Cooper pairs, and a “pseudo-gap” region where pairs form above the superfluid critical temperature. We directly observe the nonlocal nature of fermion pairing in a Hubbard lattice gas, using spin- and density-resolved imaging of ~1000 fermionic potassium-40 atoms under a bilayer microscope. Complete fermion pairing is revealed by the vanishing of global spin fluctuations with increasing attraction. In the strongly correlated regime, the fermion pair size is found to be on the order of the average interparticle spacing. Our study informs theories of pseudo-gap behavior in strongly correlated fermion systems. L ong-range Cooper pairs form in a Fermi gas for even the weakest attraction be- tween fermions. With increasing inter- action, fermion pairs become more tightly bound, as the system undergoes a smooth crossover from Bardeen-Cooper-Schrieffer (BCS) superfluidity toward a Bose-Einstein conden- sate (BEC) of molecular pairs (1–3). In the BCS limit, pair formation and the onset of super- fluidity occur at the same temperature, but in the crossover, pairs are expected to form at temperatures above the critical temperature Tc for superfluidity; the onset pair-formation temperature is usually called T (cid:1) . In this so- called “pseudo-gap” regime, the pair size should be on the order of the interparticle spacing and pairing strongly affected by many-body effects (4, 5). The character of this strongly correlated regime, situated between a Fermi liquid and a normal Bose liquid, is a matter of debate, whose resolution should affect understanding of other strongly coupled fermion systems, such as the high-Tc cuprates and twisted bi- layer graphene (6–8). The rich physics of the BEC-BCS crossover is captured by the attractive Fermi-Hubbard model, a spin-1/2 gas of fer- mions hopping on a lattice with on-site inter- actions between unlike spins (9–17). Through a particle-hole transformation, it stands in one-to-one correspondence with the repulsive Hubbard model (18, 19), which under charge doping is believed to hold the key toward un- derstanding high-temperature superconduc- tivity. The model can be realized using neutral fermionic atoms in optical lattices with tun- able interactions. Recent investigations have found spectral gaps (20), correlations between local pairs (21), and evidence for interspin cor- relations from density profiles (22). In this work, we observe the formation and spatial ordering of nonlocal fermion pairs in the pseudo-gap regime of an attractive Hubbard gas confined to two dimensions. We use bilayer quantum gas microscopy to detect the in situ location and spin of each fermion in every exper- imental shot (23–25). Access to microscopic spin and density correlations reveals the formation of nonlocal pairs, the development of long- range spatial correlations between pairs, and the interplay of pair fluctuations with this density-wave order. Department of Physics, MIT–Harvard Center for Ultracold Atoms, and Research Laboratory of Electronics, MIT, Cambridge, MA 02139, USA *Corresponding author. Email: hartke@mit.edu Fig. 1. Atom-resolved detection of an attractive Fermi-Hubbard gas. (A) Qualitative phase diagram of the attractive Fermi-Hubbard model versus on-site attraction U=t and temperature T=t at density n≈0:8 (9–13). Below a critical temperature Tc, attractive fermions form a BCS or BEC superfluid (SF). In the pseudo-gap regime between Tc and pairing temperature T(cid:1), accessed in this work (white shading), increasing attraction drives pair formation, with pairs exhibiting charge density wave (CDW) and superfluid correlations. (B) Measured doublon density d (circles) at fixed density n versus U=t, from the noninteracting limit d ¼ ðn=2Þ2 (triangles) to the fully paired limit d ¼ n=2 (squares), with representative images of the full density in ~20 × 20 site regions shown above. (C) Snapshot of full spin-and- density readout of a strongly correlated gas at U=t ¼ 8:4 4ð Þ and T=t ¼ 0:36 5ð Þ. The spin up (blue), spin down (red), and combined images (rightmost panel) are obtained via bilayer quantum gas microscopy (23, 25). Hartke et al., Science 381, 82–86 (2023) 7 July 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E Simulating the attractive Fermi-Hubbard model The phase diagram of the attractive Fermi- Hubbard model is shown in Fig. 1A as a func- tion of the attractive on-site interaction strength U , tunneling amplitude t, and temperature T (9–13). For weak attraction U ≪ t, a BCS superfluid of long-range fermion pairs forms with Tc ¼ T (cid:1) , reflecting the exponentially weak pair binding. In the opposite limit of strong attraction U ≫ t, all fermions are bound into local on-site pairs below a dissociation temperature T (cid:1) ∼ U . These pairs condense at the critical temperature of Bose-Einstein condensation Tc, proportional to the pair density np and pair tunneling rate tp ∼ t2=U. A peak of the condensation temperature Tc=t ≈ 0:2 is expected to occur at U =t ≈ 6 and density n ≈ 0:8 (9–13). Above the transition temperature, superfluid correlations compete with the formation of a checkerboard charge density wave (CDW) (21). At half-filling (den- sity n ¼ 1), this competition persists down to T ¼ 0 and prevents condensation. In this work we use a filling n ≈ 0:8, staying in a regime where the ground state is a paired superfluid (9, 15). We here experimentally realize the attractive Fermi-Hubbard model using a two-species gas of degenerate fermionic 40 K atoms trapped within an optical lattice (23, 25). The Hubbard tunneling amplitude t is controlled by the lattice depth, and the interaction strength U is tuned via the magnetic field. A bilayer quan- tum gas microscope reveals the location of each fermion. As a first measure of strong pairing in the attractive Hubbard gas, we detect the density d of doublons (doubly occupied lattice sites) for increasing interaction strength U =t across the phase diagram in Fig. 1A. At fixed density n, d increases from the noninteracting limit d ¼ ðn=2Þ2 of random encounters of unlike spins to the fully paired limit d ¼ n=2 (Fig. 1B) (17). At intermediate attraction, strong checkerboard ordering of doublons is observed, shown in Fig. 1C at U =t ¼ 8:4 4ð Þ and T =t ¼ 0:36 5ð Þ. Multiple neighboring sites containing a sin- gle spin up and spin down are present among doublons in Fig. 1C. These correlated pairs of single spins are evidence of the nonlocal na- ture of fermion pairs. The microscopic mech- anism is the virtual dissociation of a doublon into spatially separate pairing partners, with matrix element t and intermediate energy cost U, which perturbatively lowers the energy of a pair by 4t2=U. Because pairs are composed of fermions, dissociation can only occur if a nearby site does not already contain a like spin. This leads to effective nearest-neighbor repul- sive interactions between pairs (23), which in turn are the source of long-range CDW order. The presence of these delocalized pairs also demonstrates that the doublon density d is an incomplete measure of pairing. Fig. 2. Observation of nonlocal fermion pairing. (A) Experimental snapshots of the Fermi gas at U=t ¼ 0, U=t ¼ 5:8 3ð Þ, and U=t ¼ 8:4 4ð Þ (left to right), showing the formation of nonlocal pairs and on-site pairs with increasing attraction. Schematics above highlight the physics dominating spin correlations in each image, and shaded bonds suggest possible pair correlations. (B) Correlation maps h ^mi ^miþdi m ¼ n↑ (cid:3) n↓ at various U=t. (C) Total magnetization fluctuations ^miþdi c of the magnetization c (blue circles) and on-site h ^mi X → d fluctuations (black squares) versus U=t. Total fluctuations equal the product of magnetic susceptibility c temperature T via the fluctuation-dissipation theorem (23). Vanishing total spin fluctuations for U=t ≥ 6 (orange shading) indicate full pairing and vanishing c of total fluctuations at n ¼ 0:85, from T=t ¼ 0:3 to T=t ¼ 0:4 (23). The pairing temperature T(cid:1) crosses T ≈ 0:35t at U=t ≈ 2:5. Nonlocal pairing is reflected in the singlon fraction per total density s=n (upper inset), which scales as ~8t2=U2 (gray line) at large attraction. Fluctuations at U=t ¼ 5:8 3ð Þ (lower inset) extend beyond the interparticle spacing 1= bootstrapping greater than 50 images of atomic clouds with imaging loss correction (23). m. Blue shading shows quantum Monte Carlo simulations (dotted line). All data and error bars are obtained from m and ffiffiffiffiffiffiffi pn↑ p Detecting fermion pairing A true signature of pairing that accounts for these nonlocal pairs is the vanishing of total spin fluctuations. Indeed, a system in contact with a surrounding particle bath will generally display fluctuations of the total magnetiza- tion M ¼ X i h ^mii, where the magnetization m ¼ n↑ (cid:3) n↓. However, pair formation sup- presses spin fluctuations, as pairs do not contribute to M, and thus in a fully paired system the variance s2 M vanishes. This variance Hartke et al., Science 381, 82–86 (2023) 7 July 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E ^niþdi c reflect the crossover from Fig. 3. Charge density wave ordering of pairs. (A) Density correlations h^ni a noninteracting gas (left), to a fully paired gas with CDW order (center), to a weakly ordered gas of ð(cid:3)1Þdxþdy are well local pairs (right). (B) At U=t ¼ 8:4 4ð Þ, the nonlocal rectified correlations h^ni p ffiffiffiffiffiffiffi described by long-range exponential decay. Dotted line shows the interparticle spacing 1= pn↑ density response c n (red circles) and interspin response c↑↓ (blue squares) at wave vector k order parameters for CDW correlations. These susceptibilities are obtained via the fluctuation-dissipation (cid:3) theorem c ccos k →(cid:3) (cid:4) n k at U=t ¼ 0 (black) and U=t ¼ 8:4 4ð Þ (red). (D) Fluctuation thermometry: The measured (cid:3) ccos k → versus k →(cid:3) (cid:4) n k h^ni↑^niþd↓i and c↑↓ k . (C) The ^niþdi c ¼ 1=Tð ¼ 1=Tð Þ serve as (23). Inset: ¼ p; pð ^niþdi →(cid:3) (cid:4) → (cid:4) d → (cid:4) d h^ni X X (cid:4) (cid:4) c → d → d Þ Þ → → → h^ni ^niþdi c (top) and uniform compressibility @n=@m ¼ c (cid:3) → n k (cid:4) Þ ¼ 0; 0ð (middle) X density fluctuations → d are combined to directly obtain the temperature T=t (bottom). is measured locally in our quantum gas mi- croscope through the sum of connected cor- relations s2 c, where h ^mi ^miþdi h ^mi ^miþdi M = Area ð → d ¼ h ^mi ^miþdi (cid:3) h ^miih ^miþdi. Þ ¼ X c The magnetization fluctuations are directly connected to the magnetic susceptibility cm ¼ @m=@h, the response of the magnetization to a global magnetic field h, through the fluctuation- c (23). dissipation theorem cmT ¼ h ^mi ^miþdi X → d An energy gap for spin excitations, which ex- ponentially suppresses excess spins and thus X c, also exponentially suppresses h ^mi ^miþdi → d cm (26). Figure 2 reports a crossover to full fermion pairing beyond an interaction strength U =t ≈ 6 at n ¼ 0:8 1ð Þ and T =t ¼ 0:35 5ð Þ, determined by in situ observation of magnetization fluc- tuations. The reduction in fluctuations is in good agreement with theoretical predictions for these parameters (12–14). Figure 2A high- lights the physical mechanisms that determine spin fluctuations at various U =t . At vanishing interactions, Pauli exclusion separately reduces the density fluctuations of each spin, and there- Hartke et al., Science 381, 82–86 (2023) 7 July 2023 by also reduces total spin fluctuations. With increasing attraction, nonlocal pairs form in which spins are subject to a competition of Pauli exclusion and attraction, while deep in the on-site pair regime spin fluctuations reflect virtual hopping onto neighboring sites. From statistical averages over more than 50 spin configurations, as shown in Fig. 2A for each in- teraction strength, we obtain the two-dimensional (2D) magnetization correlation maps h ^mi ^miþdi c, shown in Fig. 2B. To detect pairing, Fig. 2C presents the sum of these correlation maps, the total magnetization fluctuations, which are fully suppressed beyond U =t≈6. Already at zero interactions, Pauli exclusion reduces total fluctuations by 68 5ð Þ% compared to the high- temperature expectation n 1 (cid:3) n=2 Þ. This re- flects the substantial degeneracy of the Fermi gas (T =TF ≈ 0:1, where TF is the Fermi tem- perature). Increasing attraction reduces mag- netization fluctuations further, and the fraction of unpaired spins is less than 1:5 1:8ð Þ% at X U =t ¼ 5:8 3ð Þ, where c gives the h ^mi ^miþdi → d density of unpaired spins. This full suppression is dual to the formation of a Mott insulator for repulsive interactions (18, 19, 23, 25). ð The suppression of fluctuations in Fig. 2C with increasing U =t signifies the development of an energy gap for spin excitations (26). The- ory predicts (10–14, 27, 28) a pairing tem- perature T (cid:1) ≈ 0:25U in the crossover regime (23). This predicted T (cid:1) crosses T ≈ 0:35t near U =t ≈ 2:5, explaining the near-complete sup- pression of fluctuations beyond U =t ≈ 6. The corresponding expected spin excitation gap far exceeds the two-body binding energy Eb, highlighting the many-body nature of pairing. Within this regime of full pairing, the two atoms within a pair have a finite probability of being located on separate lattice sites be- cause of quantum fluctuations, in which case they appear as isolated single atoms (called singlons). This probability is measured by the ratio of the singlon density to the total density, s=n (Fig. 2C, upper inset). The observed scaling of s=n with t2=U 2 at strong attraction is ex- pected from perturbation theory already for a Fermi-Hubbard double well (25, 29). At U =t ¼ 5:8 3ð Þ, the nonlocal portion of the pairs amounts to ~20%. The effective size of fermion pairs can be obtained as the spatial extent of nonlocal spin fluctuations. With full pairing at U =t ¼ 5:8 3ð Þ, spin fluctuations are present be- yond the single-spin interparticle spacing (Fig. 2C, lower inset), indicating that fermion pairs overlap substantially. Measuring charge correlations Characteristic for the pseudo-gap regime is a predicted strong departure from Fermi liquid behavior, in which spin and charge fluctua- tions are similar (4, 5). Having established the existence of nonlocal fermion pairs through vanishing magnetization fluctuation, we there- fore now explore charge (i.e., density) correla- tions of the gas. Whereas for weak interactions charge and spin correlations are closely asso- ciated, for stronger attraction we instead find spatial ordering into a CDW across the phase diagram of Fig. 1A. Previously, evidence for CDW order has only been observed in doublon- doublon correlations and at a fixed interaction strength (21). In Fig. 3A, beginning without interactions, we observe negative nonlocal den- sity correlations h^ni ^niþdi c for nearest-neighbor and diagonal correlations. These correlations are a simple and direct manifestation of the Pauli hole experienced by a single spin (25, 30). Indeed, here h^ni ^niþdi c is precisely equal to h^ni↑ ^nj↑i c + h^ni↓ ^nj↓i c because the measured in- terspin correlations h^ni↑ ^nj↓i c vanish (23). For increasing attraction, the Pauli hole gives way to the positive checkerboard long-range density correlations, shown versus distance in Fig. 3B at U =t ¼ 8:4 4ð Þ. Further increase in U =t re- duces the observed CDW strength, likely as a result of smaller effective repulsion between more-localized pairs (Fig. 2B). Notably, we measure a negative sum of nonlocal inter- spin correlations h^ni↑^niþd↓i c for any attractive 3 of 5 RES EARCH | R E S E A R C H A R T I C L E A Pair Fluctuation Tunneling Superfluid Charge-Density Wave _ _ Flipped SF Phase Reduced CDW Strength B Bare Correlations vs Conditioned on Singlon or C = D (cid:4) D E (cid:3) ^ di (cid:3) ^ hi the CDW strength, defined as Fig. 4. Interplay of nonlocal pairs and many-body order. (A) When local pairs in a strongly correlated attractive system virtually fluctuate into spatially separated single spins, further tunneling can disturb the underlying many-body ordered state. (B) Polaronic correlations are observed by measuring disturbances of (cid:4) ð(cid:3)1Þdxþdy, in the vicinity of a single spin. Whereas a (cid:3) diþd (cid:3) ^ ^ peak in the background CDW strength is observed near U=t ≈ 8 for various displacements d a strong reduction in CDW strength is observed after conditioning on the presence of a nearby isolated spin (right panel). Right inset shows the relative location of the single spin (purple circle) and CDW bond. (C) The relative change in CDW strength DCDW reveals the spatial extent of polaronic correlations at U=t ¼ 8:4 4ð Þ. (cid:6) →(cid:6) (cid:6)r (cid:6) from the singlon to the CDW bond. The shaded region represents the extent of DCDW decays with distance the disturbance to the background order. (D) A symmetrized 2D map of (C), with lattice sites represented by black dots and the isolated spin at center. In the CDW strength, ^ ^ hi) denotes a doublon (hole, or empty site). di ( (left panel), hiþd → c interaction (23), revealing that a single ↑ atom in total repels spin ↓ atoms on all other sites. This constitutes a strong direct signature of effective repulsion between pairs. ¼ p; pð The development and destruction of CDW or- der across the phase diagram of Fig. 1 can be cap- tured by the density response cn at wave vector → k Þ (Fig. 3C), which reflects the preva- lence of low-energy states with checkerboard order. Highlighting the power of quantum gas microscopy, this thermodynamic property can be Hartke et al., Science 381, 82–86 (2023) 7 July 2023 measured in equilibrium using the fluctuation- dissipation theorem for density perturbations, cn k , and h^ni ^niþdi (cid:3) ccos k ¼ 1=Tð (cid:3) (cid:4) → → (cid:4) d X (cid:4) Þ → → d → (cid:4) Þ ¼ 0; 0ð (cid:3) the measured uniform density compressibility cn k ¼ @n=@m (Fig. 3D) (25). This (cid:3) (cid:4) → same method allows measurement of cn k throughout the Brillouin zone (Fig. 3C, inset) and provides a model-independent measure- ment of temperatureT (Fig. 3D) (25, 31). The latter enables obtaining the magnetic susceptibility → → (cid:4) d c m from the measured spin fluctuations with- out applying a magnetic field (23, 32). As ex- pected from the phase diagram in Fig. 1, the peak in CDW order occurs near U =t ≈ 6. Also displayed are the interspin correlations, obtained (cid:4) X from c↑↓ k Þ . h^ni↑ ^niþd↓i (cid:3) ccos k ¼ 1=Tð (cid:3) (cid:4) → → d Although opposite spins are uncorrelated at U =t ¼ 0, they are seen to almost fully carry the CDW order beyond U =t ≈ 6. Because density, magnetization, and interspin correlations are re- ¼ 4h^ni↑ ^niþd↓i (cid:3) h ^mi ^miþdi lated byh^ni ^niþdi c, the relative agreement of cn and c↑↓ illustrates the strength of density order as compared to → Þ. The pronounced ¼ p; pð magnetic order at k CDW peak is also a signature of strong super- fluid correlations within the crossover regime, as CDW correlations away from half-filling serve as a lower bound for superfluid correla- tions (21, 33). c c Polaronic correlations Given simultaneous charge and spin measure- ments, we finally explore the interplay of non- local pair fluctuations and the CDW order of other pairs, revealing the existence of polar- onic correlations in the CDW order of the attractive Hubbard model. Polaronic corre- lations occur in the regime of highly nonlocal pairs, where further tunneling of a separated pair can dislocate the CDW checkerboard or flip the sign of superfluid correlations (Fig. 4A). These tunneling events prevent the virtual delocalization of other pairs across the bonds where the order has been reversed, costing an additional 4t2=U per bond in the strong- coupling limit and further confining spatially separated pairs (34). This mechanism is di- rectly complementary (18, 19, 23) to the mag- netic polaron mechanism of the repulsive Fermi-Hubbard model (35, 36), though here polaronic correlations dress the individual spins of a spatially separated fermion pair, rather than excess dopants. In Fig. 4B, we compare the CDW correlations surrounding single spins to those present in the background. We quantify the underlying CDW strength as hð^d i (cid:3) ^hiÞð^d iþd (cid:3) ^hiþdÞi ð(cid:3)1Þdxþdy c → for a given displacement d , which is positive for → any d for a gas possessing checkerboard doublon- hole correlations. This underlying CDW strength peaks near an interaction strength U =t ≈ 8. By contrast, for various U =t values, the measured CDW strength is strongly reduced after con- ditioning on the presence of a single nearby isolated spin. This reduction substantially ex- ceeds the lowest-order expectation of single- pair fluctuation events depicted in Fig. 4A, e.g., → Þ and 50% for 25% for a displacement d → Þ. The measurements directly Þ or 2; 0ð d reveal the spatial extent of these polaronic effects, captured by the relative change DCDW of conditioned to unconditioned CDW strength (Fig. 4, C and D). Virtual pair fluctuations disturb ð ¼ 0; 1 ð ¼ 1; 1 r→(cid:6) (cid:6) (cid:6) (cid:7) (cid:6) (cid:5) 4 of 5 RES EARCH | R E S E A R C H A R T I C L E the CDW order over a range of ∼2 sites at U =t ¼ 8:4 4ð Þ, with complete reduction or even re- versal of the CDW order on nearby bonds. In future work, measurements of four-point cor- relations (36) around pairs of spins will further elucidate the internal structure of these quan- tum fluctuations. Conclusions Our real-space observations of nonlocal fermion pairing and its interplay with CDW order il- lustrate the richness of the pseudo-gap regime of the attractive Hubbard model. Similar com- peting or intertwined orders are predicted for the repulsive Hubbard model. The methods can be extended further to study polaronic physics and superfluidity (37), pairing in mo- mentum space as measured in bulk 2D gases (38), to detect the p phase shift of CDW order across stripes (39), and to directly measure the BCS condensate fraction through pair correla- tions (40). RE FE RENCES AND N OT ES 1. M. Inguscio, W. Ketterle, C. Salomon, Ultracold Fermi Gases (IOS Press, 2008). 2. W. Zwerger, Ed., The BCS-BEC Crossover and the Unitary Fermi 9. R. T. Scalettar et al., Phys. Rev. Lett. 62, 1407–1410 (1989). 10. J. Singer, T. Schneider, M. Pedersen, Eur. Phys. J. B 2, 17–30 (1998). 39. T. Ying, R. T. Scalettar, R. Mondaini, Phys. Rev. B 105, 115116 (2022). 40. W. Ketterle, M. Zwierlein, Riv. Nuovo Cim. 31, 247–422 11. T. Paiva, R. R. dos Santos, R. T. Scalettar, P. J. H. Denteneer, (2008). Phys. Rev. B Condens. Matter Mater. Phys. 69, 184501 (2004). 12. T. Paiva, R. Scalettar, M. Randeria, N. Trivedi, Phys. Rev. Lett. 104, 066406 (2010). 13. R. A. Fontenele, N. C. Costa, R. R. dos Santos, T. Paiva, Phys. Rev. B 105, 184502 (2022). 41. T. Hartke, B. Oreg, C. Turnbaugh, N. Jia, M. Zwierlein, Replication Data for: Direct observation of non-local fermion pairing in an attractive Fermi-Hubbard gas, Dataverse (2023); https://doi.org/10.7910/DVN/HX07DM. 42. L. W. Cheuk et al., Phys. Rev. Lett. 114, 193001 (2015). 43. T. Hartke, B. Oreg, N. Jia, M. Zwierlein, Nature 601, 537–541 14. J. M. Singer, M. H. Pedersen, T. Schneider, H. Beck, H. Matuttis, (2022). Phys. Rev. B Condens. Matter 54, 1286–1301 (1996). 15. A. Moreo, D. J. Scalapino, Phys. Rev. Lett. 66, 946–948 44. R. Micnas, J. Ranninger, S. Robaszkiewicz, Rev. Mod. Phys. 62, 113–171 (1990). (1991). 45. J. Bardeen, L. N. Cooper, J. R. Schrieffer, Phys. Rev. 108, 16. M. Keller, W. Metzner, U. Schollwöck, Phys. Rev. B Condens. 1175–1204 (1957). Matter 60, 3499–3507 (1999). 17. J. Bauer, A. C. Hewson, N. Dupuis, Phys. Rev. B Condens. Matter Mater. Phys. 79, 214518 (2009). 18. A. F. Ho, M. A. Cazalilla, T. Giamarchi, Phys. Rev. A 79, 033620 46. L. Salasnich, F. Toigo, Phys. Rev. A 86, 023619 (2012). 47. M. Wortis, Phys. Rev. 132, 85–97 (1963). 48. K. Miyake, Prog. Theor. Phys. 69, 1794–1797 (1983). 49. C. N. Varney et al., Phys. Rev. B Condens. Matter Mater. Phys. (2009). 19. M. Gall, C. F. Chan, N. Wurz, M. Köhl, Phys. Rev. Lett. 124, 010403 (2020). 20. P. T. Brown et al., Nat. Phys. 16, 26–31 (2020). 21. D. Mitra et al., Nat. Phys. 14, 173–177 (2018). 22. C. F. Chan, M. Gall, N. Wurz, M. Köhl, Phys. Rev. Res. 2, 023210 (2020). 23. See supplementary materials. 24. J. Koepsell et al., Phys. Rev. Lett. 125, 010403 (2020). 25. T. Hartke, B. Oreg, N. Jia, M. Zwierlein, Phys. Rev. Lett. 125, 113601 (2020). 26. K. Yosida, Phys. Rev. 110, 769–770 (1958). 27. M. Randeria, J.-M. Duan, L.-Y. Shieh, Phys. Rev. B Condens. Matter 41, 327–343 (1990). 28. M. Drechsler, W. Zwerger, Ann. Phys. 504, 15–23 80, 075116 (2009). AC KNOWLED GME NTS Funding: This work was supported by the NSF through the Center for Ultracold Atoms and grant PHY-2012110, AFOSR (grant no. FA9550-16-1-0324 and MURI on molecules no. 2GG016303 PO No15323), and the Vannevar Bush Faculty Fellowship (ONR no. N00014-19-1-2631). Authors contributions: The experiment was designed by all authors. T.H., B.O., C.T., and N.J. collected and analyzed the data. All authors contributed to the manuscript. Competing interests: The authors declare no competing financial interests. Data and materials availability: The data and code for this manuscript are available on Dataverse (41). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. sciencemag.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.ade4245 Supplementary Text Figs. S1 to S9 References (42–49) Submitted 15 August 2022; accepted 26 May 2023 10.1126/science.ade4245 Gas (Springer, 2012). (1992). 3. M. Randeria, E. Taylor, Annu. Rev. Condens. Matter Phys. 5, 209–232 (2014). 4. M. Randeria, N. Trivedi, A. Moreo, R. T. Scalettar, Phys. Rev. Lett. 69, 2001–2004 (1992). 5. N. Trivedi, M. Randeria, Phys. Rev. Lett. 75, 312–315 (1995). 6. P. A. Lee, N. Nagaosa, X.-G. Wen, Rev. Mod. Phys. 78, 17–85 (2006). 7. Y. Cao et al., Nature 556, 43–50 (2018). 8. Q. Chen, J. Stajic, S. Tan, K. Levin, Phys. Rep. 412, 1–88 (2005). 29. S. Murmann et al., Phys. Rev. Lett. 114, 080402 (2015). 30. L. W. Cheuk et al., Science 353, 1260–1264 (2016). 31. Q. Zhou, T.-L. Ho, Phys. Rev. Lett. 106, 225301 (2011). 32. C. Sanner et al., Phys. Rev. Lett. 106, 010402 (2011). 33. P. T. Brown et al., Science 357, 1385–1388 (2017). 34. F. Grusdt et al., Phys. Rev. X 8, 011046 (2018). 35. J. Koepsell et al., Nature 572, 358–362 (2019). 36. J. Koepsell et al., Science 374, 82–86 (2021). 37. J. K. Chin et al., Nature 443, 961–964 (2006). 38. M. Holten et al., Nature 606, 287–291 (2022). Hartke et al., Science 381, 82–86 (2023) 7 July 2023 5 of 5
10.1126_science.adf8999
resonant electric dipole-dipole coupling for coherent exchange of rotational angular mo- mentum between pairs of laser-cooled CaF molecules individually trapped in optical twee- zers. With the ability to tune the angle of the molecular quantization axis in the lab frame, our system realizes both ferromagnetic and antiferromagnetic couplings in a quantum XY spin-exchange model by effectively encoding a spin-1 2 system into molecular rotational states. Using this spin-exchange Hamiltonian, we per- formed iSWAP two-qubit gate operations. An iSWAP gate, when sequentially combined with single-qubit operations, can generate bipartite entanglement deterministically and form a universal set of quantum gates, which is the essential resource for all quantum information applications (12, 43). We found that the fidelity of generated Bell states is presently limited by the thermal motion of molecules within the tweezer traps. Finally, we demonstrate the use of an interleaved dual tweezer-array system, which allows for robust single-site address- ability and fast gate operations between mo- lecules. This, combined with expected improved molecular cooling (44), has the potential to greatly increase the fidelity of two-qubit ope- rations in this system. This work is parallel and concurrent with that of another study (45). Initial-state preparation of CaF molecules in a 1D optical tweezer array We began the experiment by loading CaF mol- ecules from a cryogenic buffer gas beam source (46) into a radio-frequency magneto-optical trap (14). Next, we loaded the molecules into a 1D optical lattice using L-enhanced gray molasses D RES EARCH R E S E A R C H A R T I C L E ◥ QUANTUM INFORMATION Dipolar spin-exchange and entanglement between molecules in an optical tweezer array Yicheng Bao1,2*, Scarlett S. Yu1,2, Loïc Anderegg1,2, Eunmi Chae3, Wolfgang Ketterle2,4, Kang-Kuen Ni1,2,5, John M. Doyle1,2 Ultracold polar molecules are promising candidate qubits for quantum computing and quantum simulations. Their long-lived molecular rotational states form robust qubits, and the long-range dipolar interaction between molecules provides quantum entanglement. In this work, we demonstrate dipolar spin-exchange interactions between single calcium monofluoride (CaF) molecules trapped in an optical tweezer array. We realized the spin-1 2 system into the rotational states of the molecules and used it to generate a Bell state through an iSWAP operation. Conditioned on the verified existence of molecules in both tweezers at the end of the measurement, we obtained a Bell state fidelity of 0.89(6). Using interleaved tweezer arrays, we demonstrate single-site molecular addressability. 2 quantum XY model by encoding an effective spin-1 (28, 31, 39, 40). In this work, we report the entanglement of pairs of molecules—the crit- ical ingredient in quantum computing and simulation based on molecules—by leverag- ing long molecular coherence and intrinsic dipolar interactions with individual parti- cle addressability. A key step toward using ultracold polar mol- ecules for quantum simulations and multi- particle quantum gates is the generation of coherent dipole-dipole couplings between molecules (12, 41), which has been shown in sparsely filled three-dimensional (3D) lattices (27) and 2D layers (42) and in a molecular quantum gas microscope (33). Here, we used A B C Q uantum entanglement is one of the key ingredients that fundamentally distin- guishes quantum mechanics from clas- sical mechanics (1). It is considered an essential resource for quantum informa- tion processing but remains generally challeng- ing to create experimentally. One mechanism of realizing quantum entanglement between particles is using the electric dipole-dipole interaction, as demonstrated in systems such as Rydberg atoms (2) and silicon quantum dots (3). Polar molecules possess permanent, long-range, and spatially anisotropic electric dipoles whose interaction could be harnessed for high-fidelity entanglement generation. There- fore, they have been proposed as a powerful platform for realizing quantum simulations of strongly interacting many-body dynamics (4–7) and for scalable quantum computing (8–12). Ultracold molecules can be produced through either direct laser cooling (13–17), assembly of individual ultracold atoms (18–24), or optoelectrical Sisyphus cooling (25, 26). Cur- rent experimental progress has established the capability of preparing and manipulating ultra- cold molecules with high fidelity (27–33). In particular, rearrangeable tweezer arrays offer an attractive quantum platform owing to their scalability and potential for single-site address- ability (24, 34–38). Toward this aim, molecules have been demonstrated to have long qubit (rotational and hyperfine state) coherence times, which are substantially longer than the predicted molecule-molecule dipolar gate times 1Department of Physics, Harvard University, Cambridge, MA 02138, USA. 2Harvard-MIT Center for Ultracold Atoms, Cambridge, MA 02138, USA. 3Department of Physics, Korea University, Seongbuk-gu, Seoul 02841, South Korea. 4Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. 5Department of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138, USA. *Corresponding author. Email: bao@g.harvard.edu Fig. 1. Experimental setup and relevant energy levels. (A) The image of averaged fluorescence from a 20-site optical tweezer array of CaF molecules. (B) An illustration of the optical tweezers and the coordinate system used in this work. For visual clarity, only four sites of the realized 20-site array are depicted. → (C) An illustration showing the relative angle between the applied bias magnetic field B → and the intermolecular vector r hyperfine (F) states in the ground electronic state of CaF. . The tweezer light propagates along Z. (D) Relevant rotational (N) and (in the XY plane) Bao et al., Science 382, 1138–1143 (2023) 8 December 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E cooling on the electronic X 2Sþ–A2P1=2 tran- sition (47, 48) and optically transported them into a glass cell using a focus-tunable moving lattice, as characterized previously (49). We used a microscope objective with 0.6 nu- merical aperture to project a 1D optical tweezer array in the glass cell, as well as collect fluo- rescence from the molecules during L-imaging (47). The array is formed by passing single- frequency 776-nm laser light through a shear- mode acousto-optic deflector (AOD) driven with a multitone waveform (37), which is generated by a high-speed arbitrary waveform generator (AWG). This allows for control of the position of individual optical tweezer sites as well as trap depth. We started with a 20-site array formed by a single AOD (Fig. 1A). We loaded the molecules into the tweezers by overlapping the array with a transported mol- ecular cloud that is held in an optical dipole trap in the presence of the L-cooling light. We ob- served an average probability of 35% for loading a single site with a single molecule. After loading the array, we applied a L-imaging pulse and collected fluorescence onto a camera to identify the tweezer sites that are loaded with single molecules in the ground electronic and vibra- tional state (or empty) with a detection fidelity of 98(1)% and nondestructive detection fidelity of 94(1)% (37, 47). Because the L-cooling tech- nique relies on closed photon cycling between the X, N = 1 and A, J = 1/2 manifolds (N is the rotational angular momentum, and J is the total angular momentum, excluding nuclear spin), only molecules in the X, N = 1 rotational manifold can be loaded into the tweezers and detected in this phase of the experiment. Imaging distributes the population of mol- ecules over all 12 hyperfine states. To prepare the molecules in a single quantum state, we used a combination of optical pumping and microwave transfer. Using short X−A laser pulses resonant with all hyperfine levels in the X, v = 0, N = 1 manifold except the X ; N ¼ 1; F ¼ 0; mF ¼ 0 i state (v is the vib- j rational quantum number; F is the total angu- lar momentum, including nuclear spin; and mF is the magnetic quantum number of F), we pumped most of the molecular population into the latter state. We then linearly ramped down the trap depth of the tweezers in 5 ms and applied a bias magnetic field of ≈3.2 G. A subsequent microwave p-pulse transfers the population from X ; N ¼ 1; F ¼ 0; mF ¼ 0 i to the X ; N ¼ 0; F ¼ 1; mF ¼ 0 i state (Fig. 1B). Finally, we applied an X−A laser pulse that contains frequencies to drive all the hyperfine components in order to remove any molec- j j j ular population left over in the X, N = 1 man- ifold. In the end, all the remaining molecules in the array were in the X ; N ¼ 0; F ¼ 1; mF ¼ 0i state, initializing the qubits and ef- fectively encoding a spin- 1 2 model in the sub- space spanned by the X ; N ¼ 0; F ¼ 1; mF ¼ j 0i ≡ ↑j i and X ; N ¼ 0; F ¼ 1; mF ¼ 0 i ≡ ↓j i states (Fig. 1B). The total state preparation and detection efficiency is 60(2)%, which is limited by the residual population in other hyperfine states within X, N = 1 and imperfect imaging fidelity. This state preparation effici- ency can be improved by an optimized optical pumping scheme (45). j Single-molecule rotational coherence time To observe high-fidelity dipolar spin-exchange interactions, a long rotational coherence time comparable to the timescale of the dipolar interaction is required. With a sample of mol- ecules at finite temperature in optical tweezers, it is important to control the differential ac Stark shift broadening caused by the molecule’s thermal motion in the tweezer trap (28, 50–52). To suppress this broadening, the tweezer laser is linearly polarized at a “magic” angle relative to the quantization axis that is defined by the applied bias magnetic field, as detailed previous- ly (31) (Fig. 1A). To maximize the single-qubit A XY8 Block X π/2 X π Y π X π Y π Y π X π Y π X π X π/2 B 0.4 ) . u . a ( t s a r t n o C 0.3 0.2 0.1 0 0 D 80 60 40 20 ) z H ( h / J 0 1.5 Bare Ramsey = 3.6(6) ms C Spin Echo = 33(5) ms C XY8 C = 630(90) ms 100 200 300 Time (ms) 400 500 600 2 Tweezer Spacing ( m) 2.5 C n o i t c a r F n o i t c a r F n o i t c a r F n o i t c a r F 0.4 0.2 0 0.4 0.2 0 0.4 0.2 0 0.4 0.2 0 0 R=1.5 m, D =35(10) ms, T=35(2) ms R=1.75 m, D =75(17) ms, T=53(2) ms R=2.0 m, D =96(21) ms, T=77(2) ms R=2.25 m, D =159(29) ms, T=105(2) ms 50 100 150 200 250 Time (ms) Fig. 2. Rotational coherence and dependence of dipolar interactions on tweezer spacing. (A) The Ramsey sequence with XY8 dynamical decoupling used in this work. Xp/2 represents a p 2-pulse, which rotates the quantum state around the x axis of the Bloch sphere by 90°. Similarly, Xp (Yp) represents a 180° rotation around the x axis (y axis) on the Bloch sphere. At the top, “×N” means that the block of XY8 pulses is repeated multiple times during the evolution time. (B) Measured single-molecule rotational coherence time between ↑j i (N = 1) and ↓j i (N = 0) using Ramsey, single–p-pulse spin echo, and the XY8 dynamical decoupling sequences. The coherence time is fitted as a Gaussian 1/e decay time constant. (C) Dipolar spin-exchange oscillation at various tweezer spacings, with fitted decay-time constant and oscillation period shown in the legends. →j. (D) Dipolar spin-exchange interaction strength J versus the tweezer spacing jR The dashed orange line is the theoretical prediction of J at zero temperature. The solid blue line is the simulated result of J with the thermal motion of the molecules taken into account (59). In (B) and (C), error bars represent one standard deviation of uncertainty. Bao et al., Science 382, 1138–1143 (2023) 8 December 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E n o i t c a r F 0.5 0.4 0.3 0.2 0.1 0 0 = 54.7° = 0°, D = 90°, D = 30(7) ms, T = 46(3) ms = 96(21) ms, T = 77(2) ms 50 100 150 200 250 Time (ms) Fig. 3. Dipolar spin-exchange interaction at different angles. Shown are the data for q = 0°, 54.7°, and 90°. The q = 0° and q = 90° data are fitted to an exponentially decaying sinusoidal model, with decay-time constant tD and oscillation cycle time T shown in solid lines. Error bars represent one standard deviation of uncertainty. coherence time, it is beneficial to adiabatically lower the trap depth as much as possible without unduly spilling molecules from the trap. However, this is not optimal for maximizing the number of observed dipolar spin-exchange oscillation cycles. At finite temperature, the instantaneous dipolar interaction strength fluc- tuates owing to thermal motion, which becomes more prominent in a shallow trap. This is the main mechanism of dephasing. To mitigate this issue, we confined the molecules more tightly by operating the tweezer at a higher trap light intensity for which the magic angle is close to 90°. The work described in this man- uscript was performed under these conditions. We measured the rotational qubit coherence time under conditions of three different micro- wave pulse sequences used to drive transitions between ↑j i and ↓j i and observed the pop- ulation in ↑j i. With a Ramsey sequence, we observed a single-qubit coherence time of tc = 3.6(6) ms. By adding a single p spin-echo pulse, tc is extended to 33(5) ms. We used active mag- netic field cancellation to remove long-term drift on the order of 10 mG, but the system we used does not remove magnetic field fluc- tuation within a power line cycle time of less than 1/(60 Hz). This limits the effective interval between spin-echo pulses to an integer multiple of 60 Hz, making it difficult to measure fast dipolar oscillations with only a single p-pulse. Instead, we used dynamical decoupling schemes to preserve the qubit coherence. This technique is used in a variety of quantum information systems (53, 54), including molecular systems (27, 30, 45). We chose the XY8 dynamical de- coupling sequence (Fig. 2A) with a cycle length of 1.6 ms (55), which is much shorter than 1/ (60 Hz), and achieved a coherence time of tc = 630(90) ms. This dynamical decoupling was used for all of our measurements except where noted. Figure 2B shows the measured contrast versus time for single-particle oscil- lations between ↑j i and ↓j i, from which tc is determined. Coherent dipolar spin-exchange interaction The dipolar spin-exchange interaction Hamiltonian (41) is (cid:1) J Hdip ¼ þ þ (cid:2) (cid:2)^S þ ^S ^S ^S 2 1 2 1 (cid:3) x y y^S x þ ^S ^S 2 1 2 1 (cid:1) 2 ¼ J ^S (cid:3) ð1Þ þ (cid:2) , ^Si x, ^Si where, respectively, ^Si (^Si y) is the spin- 1 2 raising (lowering, Pauli-X, Pauli-Y) operator for molecule number i in a tweezer pair. J is the dipolar interaction strength that can be further expressed as ð2Þ J ¼ (cid:5) (cid:4) d2 4pD0r3 1 (cid:2) 3cos2q where d is the transition dipole moment be- tween the ↑j i and ↓j i state (d ≈ 1 Debye), D0 is the vacuum permittivity, r is the intermolecular spacing, and q is the angle between the quan- tization axis and the intermolecular axis direc- tion. Apart from molecular systems (27, 30, 33), this XY spin Hamiltonian was previously stu- died with Rydberg atoms in optical tweezers (56) and atoms in optical lattices (57, 58). We set the bias magnetic field perpendicu- lar to both the k-vector of the tweezer light and the direction of the 1D tweezer array. The polarization of the linearly polarized tweezer light is rotated to the tweezer array direction (i.e., a magic angle close to 90°). This config- uration provides the largest “magic trap depth” (the tightest confinement) at a given magnetic field. We first prepared pairs of tweezers spaced →j∼5 mm, with each site loaded with a at jR single molecule prepared in the ↓j i state (or empty), yielding the ↓↓j i state. At this sepa- ration, the dipolar interaction strength J is negligible (J/h < 3 Hz, where h is Planck’s constant). In a time period of ~1 ms, we then i ð ð p p Þ= Þ= ffiffiffi 2 i þ ↑↓j moved the even numbered sites toward the odd numbered sites by sweeping the AOD frequency tones of the even sites, which reduces the sep- aration and has the effect of increasing the dipolar interaction strength. We then applied a p 2-pulse to prepare both molecules in the super- position state ↑j i þ ↓j i Þ= . Then, we applied XY8 dynamical decoupling sequence of micro- wave pulses. Under the time evolution of dipolar spin-exchange Hamiltonian (Hdip), a relative p ffiffiffi phase accumulated between ↓↑j 2 ffiffiffi and ↑↑j ð i þ ↓↓j i . (Because the XY8 se- 2 quence only contains p-pulses, it will not affect the phase accumulation during the evolution under Hdip.) After a wait time, we applied an- other p 2-pulse and then moved the molecules apart. To read out the final qubit state, we used a second L-imaging pulse to project the system to ↑↑j i. We selected the data where both sites in a tweezer pair are initially loaded with single molecules. Our measurement yielded the probability P↑↑ of detecting both molecules in the ↑j i state. To summarize, starting from the initial state ↓↓j i, a microwave pulse sequence creates a final state that evolves in i h (cid:1) (cid:1) time as y tð Þ ¼ 1 1 þ e(cid:2)i Jt 1 (cid:2) e(cid:2)i Jt ↑↑j i , 2 resulting in the probability P↑↑ ¼ cos2 Jt 4ħ ¼ (cid:5) (cid:4) 2 1 þ cos Jt 1 , which oscillates at an angular 2ħ frequency of wJ ¼ J 2ħ , where ħ is reduced Planck’s constant. i (cid:2) ↓↓j (cid:3) (cid:3) 2ħ 2ħ At smaller tweezer spacings, we observed increased wJ because of the stronger dipolar interaction between the two molecules (Fig. 2C). By fitting the data to an exponentially decaying sinusoidal model, we extracted the dipolar oscil- lation cycle period T and contrast decay time constant tD. The dipolar spin-exchange strength J at different tweezer spacings can then be cal- culated from T (Fig. 2D). We found that the measured J is slightly smaller than the theo- retical prediction and deviates more as the spacing decreases. This can be explained by the finite temperature of the molecules causing (cid:7) the effective intermolecular spacing r→(cid:7) (cid:7) to be →j. We used larger than the tweezer spacing jR Monte Carlo simulations to describe the behav- ior of the molecules in the tweezer, including the thermal motion of the molecules, and show in Fig. 2D that the simulated results agree with the experimental data. Additionally, the sim- ulation captures that the thermal motion re- duces the observed number of coherent dipolar oscillations as tweezer spacing is decreased (59). (cid:7) Anisotropy of the dipolar interaction The general dipole-dipole Hamiltonian described in Eqs. 1 and 2 is inherently anisotropic owing to the q-dependent term. We studied the effect of the anisotropy of spin exchange experimentally by varying the angle q and measuring the cor- responding dipolar interaction strength. The angle q is varied by rotating the quantization axis relative to the line between the centers of the tweezers. The quantization axis is set by Bao et al., Science 382, 1138–1143 (2023) 8 December 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E the applied bias magnetic field and can be changed by tuning the current through two pairs of magnetic field coils while simultaneously ro- tating the tweezer light polarization to maintain the same magic angle. In Fig. 3, we show P↑↑ at three characteristic angles q = 0°, 54.7°, and 90°, all taken with a →j ¼ 2 mm tweezer spacing. We observed di- jR polar spin-exchange oscillations in both the q = 0° and q = 90° configurations, with the oscil- lation at q = 0° being at twice the frequency as that at q = 90°. No clear oscillation is observed for the q = 54.7° configuration, which was as ex- pected because the dipolar interaction averages to zero at this angle. These results agree with the absolute value of the magnitude of the anis- otropic term in each configuration. We also ob- served a larger number of oscillation cycles at q = 90° than at q = 0°. Our Monte Carlo sim- ulation indicated that thermal motion dephas- ing is the dominant cause. At q = 0°, the larger motional wave function spread in the more- weakly trapped axial direction of the optical tweezers results in a large fluctuation of the instantaneous value of q. The q = 90° config- uration is less affected because of the tight confinement of the optical tweezers in the ra- dial direction. Fidelity of created Bell states Dipolar spin-exchange can be used in a two- qubit iSWAP gate to generate entanglement between molecules in neighboring tweezers (12). At a dipolar spin-exchange interaction time of t = T/4 = 19.2 ms, where T is the oscillation period of the spin-exchange oscilla- tion, the system has evolved into a maximally entangled state known as a Bell state. To test the fidelity of the Bell state that was generated in our system, we applied a third p 2-pulse (see Fig. 4 legend) around a variable rotation axis on the Bloch sphere (angled f relative to the x axis on the equatorial plane) (Fig. 4A). By vary- ing f and measuring the survival probability of all four possible final-state outcomes, one can construct the parity quantity P (43, 60): D ¼ ^S (cid:8) (cid:9) P ¼ ^P E z z ^S 2 1 ¼ P↑↑ þ P↓↓ (cid:2) P↑↓ (cid:2) P↓↑ ð3Þ Here,hi denotes averaging over all occurrences where both sites in a tweezer pair are initially loaded with single molecules, and ^Si represents the Pauli-Z operator on the number i molecule in a pair. z Starting from a Bell state, this sequence will result in a 4p oscillation in P as f is varied from 0 to 2p (61). In Fig. 4B, P is displayed for both the q = 0° and q = 90° configu- rations. Extracting the contrast of the oscil- lation for the q = 90° case, we measured a Bell state fidelity of F ¼ 0:32 2ð Þ and a state prep- aration and measurement (SPAM)–corrected fidelity of F SPAM ¼ 0:89 7ð Þ. The phase of the parity oscillation reveals the sign of the aniso- tropic term (1 – 3cos2q), also seen previously (45). Our data show that the q = 0° configuration leads to a negative J, which corresponds to a ferromagnetic interaction, and the q = 90° con- figuration leads to a positive J, which corre- sponds to an antiferromagnetic interaction. j ð F SPAM is measured to be significantly higher than F, with imperfect state preparation and detection being the surmised cause. In detail, during initial-state preparation, molecules that fail to be prepared in the desired N ¼ 0; F ¼ 1; mF ¼ 0i ↓j i Þ state are intentionally removed by a resonant laser pulse, resulting in an empty trap or traps. During the final readout, if the molecule is in the ↓j i state or the trap is simply empty, both would appear dark. We used a p-pulse and additional imaging pulse (Fig. 4C) to distinguish between these two cases. If the molecule is in the ↓j i state, it will be trans- ferred to the ↑j i state and detected, whereas an empty trap will remain dark. This information is used to exclude the cases of empty traps during the final readout, which improves the contrast of the parity oscillation (Fig. 4D). The resulting Bell state fidelity corrected for measurement error is determined to be F ex ¼ 0:89 6ð Þ, which is higher than the threshold of F th ¼ 0:5 (60), showing, under these conditions, the conditional preparation of entanglement of two molecules in a tweezer pair. The evolution of the system to create a Bell state is an iSWAP operation, and, although we have not fully characterized the system as an iSWAP gate, the Bell state fidelity indicates how it would perform. XY8 Block X π/2 π/2 #2 X π #3 90° Parity, Empty Traps Excluded, A = 0.81(6) A B X π/2 1 0.5 0 -0.5 -1 y t i r a P XY8 Block X π/2 π/2 #2 0° Parity, A = -0.31(7) 90° Parity, A = 0.31(4) C D y t i r a P X π/2 1 0.5 0 -0.5 -1 0 45 90 135 180 (Degrees) 225 270 315 360 0 45 90 135 180 (Degrees) 225 270 315 360 Fig. 4. Parity measurements showing the creation of Bell state pairs. (A) The microwave pulse and detection sequence used in the parity oscillation p measurement. The XY8 block is the same as that in Fig. 2A. fp/2 denotes a 2 -pulse with a microwave phase shifted by f relative to the first Xp/2 pulse in the sequence, effectively rotating the quantum state around an axis that is angled f relative to the x axis on the Bloch sphere. The pulse in yellow with a camera symbol represents imaging of the molecules. Note that the first imaging pulse for tweezer-loading identification, as well as the microwave pulses for initial-state preparation, are applied before this sequence and not shown in this figure. (B) Parity oscillation at q = 0° and q = 90°. A is the fitted parity oscillation amplitude. (C) With the addition of a p-pulse and a third imaging step to the sequence shown in (A), molecules are verified to be present throughout the entanglement and readout process. (D) Parity oscillation for q = 90°, with empty traps excluded and corrected for measurement error, using the sequence depicted in (C). In (B) and (D), error bars represent one standard deviation of uncertainty. Bao et al., Science 382, 1138–1143 (2023) 8 December 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E A n o i t c a r F 0.8 0.6 0.4 0.2 0 -0.2 0 | | , T = 21(2) ms , T = 22(2) ms B 0.6 0.5 0.4 0.3 0.2 0.1 n o i t c a r F Odd Sites Even Sites 5 10 15 Time ( s) 20 25 30 0 0 10 20 30 40 Time (ms) 50 60 70 Fig. 5. Single-site–resolved state preparation. (A) Rabi oscillation after preparation of the molecular pair in ↓↑j ↑↓j i state at q = 90°. Shown here is the outcome of ↓↑j i. In (A) and (B), error bars represent one standard deviation of uncertainty. i and ↑↓j i. (B) Dipolar spin-exchange oscillation with an initial Toward arbitrary initial-state preparation Motivated by the desire to perform robust single-site addressing, instead of using a single AOD to generate all sites in the array, we switched to using one AOD to generate the odd numbered sites and another AOD to generate the even numbered sites. This allowed for conve- nient independent trap-depth control over each molecule in a pair and uniformity across the array. Additionally, by offsetting the frequency of the tweezer light of the even and odd sites, molecules can be moved in close proximity without experiencing the heating that can arise in a single AOD system (34, 38). For a given trap depth, the differential ac Stark shift results in molecules in even numbered sites being away from the resonance of the ↑j i→ ↓j i microwave transition, therefore allowing separate micro- wave addressing of the odd sites. By applying a microwave p-pulse when odd sites are detuned away, we could prepare an antiferromagnetic i. Under Hdip, an initial state ↑↓j initial state ↑↓j i (cid:7) (cid:7) evolves as y tð Þ ¼ cos Jt 2ħ ↑↓i (cid:2) isin Jt j 2ħ ↓↑i , and (cid:5) (cid:4) 2 1 þ cos Jt P↓↑ ¼ cos2 Jt 2ħ ¼ 1 will thus oscillate ħ at an angular frequency of J ħ. To demonstrate individual addressing, we first prepared the molecules (both in the even and odd sites) in initial state ↓↓j i . We then adiabatically ramped the trap depth of the odd sites to seven times that of the even sites, so as to detune the transition of the molecules in the odd sites out of resonance. The microwave p-pulse then only transfers the molecules in the even sites from ↓j i to ↑j i. This creates an antiferromagnetic state ↓↑j i. By then applying a microwave pulse with variable length of time and detecting molecules in the ↑j i state, we ob- served Rabi oscillations in both even and odd sites with opposite phase (Fig. 5A). To observe dipolar spin exchange, we moved the tweezers →j ¼ 2 mm). As with the to a smaller spacing (jR single AOD system, we applied the XY8 dynam- ical decoupling pulses and then separated the pairs for detection. The resultant outcome pro- babilities are shown in Fig. 5B, with a clear dis- play of spin exchange. Conclusions and outlook We observed dipolar spin-exchange interac- tions and created Bell-state entangled pairs with single CaF molecules trapped in optical tweezers. We studied the dipolar interaction and entanglement by tuning the spacing of the optical tweezers and the angle of the electric dipole quantization axis. By applying detection at the end of the entanglement sequence to include only cases where two molecules are present, we determined a Bell state fidelity of F ex ¼ 0:89 6ð Þ through a parity oscillation measurement. Parity measurements confirm that the interaction in this system can be tuned between ferromagnetic and antiferromagnetic. The coherence time of dipolar interactions and the single-molecule rotational coherence time are both limited by the finite tempera- ture of the molecules. Implementation of fur- ther cooling using other techniques, for example, Raman sideband cooling (44), would substan- tially reduce motional dephasing, extend the single-qubit rotational coherence time (31), and thus increase the two-qubit gate fidelity (12). The approach presented in this work can be extended to ultracold polyatomic molecules, which have robust parity doublet states that give rise to an advantageous Stark level struc- ture (62, 63). RE FERENCES AND NOTES J. Preskill, Cal. Inst. Technol. 16, 1–8 (1998). 1. 2. D. Barredo et al., Phys. Rev. Lett. 114, 113002 (2015). 3. D. M. Zajac et al., Science 359, 439–442 (2018). 4. A. Micheli, G. Brennen, P. Zoller, Nat. Phys. 2, 341–347 (2006). 5. B. Gadway, B. Yan, J. Phys. At. Mol. Opt. Phys. 49, 152002 (2016). 6. M. A. Baranov, Phys. Rep. 464, 71–111 (2008). 7. M. Wall, K. Hazzard, A. M. Rey, From Atomic to Mesoscale: The Role of Quantum Coherence in Systems of Various Complexities (World Scientific, 2015), pp. 3–37. 8. D. DeMille, Phys. Rev. Lett. 88, 067901 (2002). 9. S. Yelin, K. Kirby, R. Côté, Phys. Rev. A 74, 050301 (2006). 10. M. Karra, K. Sharma, B. Friedrich, S. Kais, D. Herschbach, J. Chem. Phys. 144, 094301 (2016). 11. R. Sawant et al., New J. Phys. 22, 013027 (2020). 12. K.-K. Ni, T. Rosenband, D. D. Grimes, Chem. Sci. 9, 6830–6838 (2018). 13. J. F. Barry, D. J. McCarron, E. B. Norrgard, M. H. Steinecker, D. DeMille, Nature 512, 286–289 (2014). 14. L. Anderegg et al., Phys. Rev. Lett. 119, 103201 (2017). 15. S. Truppe et al., Nat. Phys. 13, 1173–1176 (2017). 16. A. L. Collopy et al., Phys. Rev. Lett. 121, 213201 (2018). 17. N. B. Vilas et al., Nature 606, 70–74 (2022). 18. K.-K. Ni et al., Science 322, 231–235 (2008). 19. P. K. Molony et al., Phys. Rev. Lett. 113, 255301 (2014). 20. T. Takekoshi et al., Phys. Rev. Lett. 113, 205301 (2014). 21. J. W. Park, S. A. Will, M. W. Zwierlein, Phys. Rev. Lett. 114, 205302 (2015). 22. M. Guo et al., Phys. Rev. Lett. 116, 205303 (2016). 23. T. M. Rvachov et al., Phys. Rev. Lett. 119, 143001 (2017). 24. L. R. Liu et al., Science 360, 900–903 (2018). 25. M. Zeppenfeld et al., Nature 491, 570–573 (2012). 26. A. Prehn, M. Ibrügger, R. Glöckner, G. Rempe, M. Zeppenfeld, Phys. Rev. Lett. 116, 063005 (2016). 27. B. Yan et al., Nature 501, 521–525 (2013). 28. F. Seeßelberg et al., Phys. Rev. Lett. 121, 253401 (2018). 29. H. J. Williams et al., Phys. Rev. Lett. 120, 163201 (2018). 30. J.-R. Li et al., Nature 614, 70–74 (2023). 31. S. Burchesky et al., Phys. Rev. Lett. 127, 123202 (2021). 32. W. B. Cairncross et al., Phys. Rev. Lett. 126, 123402 (2021). 33. L. Christakis et al., Nature 614, 64–69 (2023). 34. M. Endres et al., Science 354, 1024–1027 (2016). 35. D. Barredo, S. de Léséleuc, V. Lienhard, T. Lahaye, A. Browaeys, Science 354, 1021–1023 (2016). 36. H. Bernien et al., Nature 551, 579–584 (2017). 37. L. Anderegg et al., Science 365, 1156–1158 (2019). 38. J. T. Zhang et al., Quantum Sci. Technol. 7, 035006 (2022). 39. J. W. Park, Z. Z. Yan, H. Loh, S. A. Will, M. W. Zwierlein, Science 357, 372–375 (2017). 40. P. D. Gregory, J. A. Blackmore, S. L. Bromley, J. M. Hutson, S. L. Cornish, Nat. Phys. 17, 1149–1153 (2021). 41. A. V. Gorshkov et al., Phys. Rev. Lett. 107, 115301 (2011). 42. W. G. Tobias et al., Science 375, 1299–1303 (2022). 43. Q. Turchette et al., Phys. Rev. Lett. 81, 3631–3634 (1998). 44. L. Caldwell, M. Tarbutt, Phys. Rev. Res. 2, 013251 (2020). 45. C. M. Holland, Y. Lu, L. W. Cheuk, arXiv:2210.06309 [cond-mat. quant-gas] (2022). 46. N. R. Hutzler, H.-I. Lu, J. M. Doyle, Chem. Rev. 112, 4803–4827 (2012). 47. L. W. Cheuk et al., Phys. Rev. Lett. 121, 083201 (2018). 48. L. Anderegg et al., Nat. Phys. 14, 890–893 (2018). 49. Y. Bao et al., New J. Phys. 24, 093028 (2022). 50. B. Neyenhuis et al., Phys. Rev. Lett. 109, 230403 (2012). 51. E. Chae, Phys. Chem. Chem. Phys. 23, 1215–1220 (2021). 52. J. Lin, J. He, X. Ye, D. Wang, Phys. Rev. A 103, 023332 (2021). Bao et al., Science 382, 1138–1143 (2023) 8 December 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 53. T. Gullion, D. B. Baker, M. S. Conradi, J. Magn. Reson. 89, 479 65. Y. Bao et al., Data for Dipolar spin-exchange and entanglement (1990). 54. J. Du et al., Nature 461, 1265–1268 (2009). 55. The XY8 sequence effectively acts as a band-pass filter with, in our case, its pass-band centered at 10 kHz (64), which is higher than the dominant magnetic field noise spectrum range in our system. 56. S. de Léséleuc, D. Barredo, V. Lienhard, A. Browaeys, T. Lahaye, Phys. Rev. Lett. 119, 053202 (2017). 57. A. de Paz et al., Phys. Rev. Lett. 111, 185305 (2013). 58. P. N. Jepsen et al., Nature 588, 403–407 (2020). 59. See fig. S2 in the supplementary materials. 60. C. A. Sackett et al., Nature 404, 256–259 (2000). 61. L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, P. Treutlein, Rev. Mod. Phys. 90, 035005 (2018). 62. P. Yu, L. W. Cheuk, I. Kozyryev, J. M. Doyle, New J. Phys. 21, 093049 (2019). 63. C. Hallas et al., Phys. Rev. Lett. 130, 153202 (2023). 64. J. Choi et al., Phys. Rev. X 10, 031002 (2020). between molecules in an optical tweezer array. Zenodo (2023); https://doi.org/10.5281/zenodo.8404386. ACKN OWLED GMEN TS We thank M. D. Lukin and N. Y. Yao for carefully reading the manuscript. Funding: This material is based upon work supported by the US Department of Energy, Office of Science, National Quantum Information Science Research Centers, Quantum Systems Accelerator. Additional support is acknowledged from the Air Force Office of Scientific Research, the Asian Office of Aerospace Research and Development, the Army Research office, and the National Science Foundation (NSF). S.S.Y. acknowledges support from the NSF Graduate Research Fellowships Program. L.A. and S.S.Y. acknowledge support from the Harvard Quantum Initiative (HQI). E.C. acknowledges support from the National Research Foundation of Korea (2020R1A4A1018015, 2021M3H3A1085299, 2022M3E4A1077340, and 2022M3C1C8097622). Author contributions: Y.B., S.S.Y., L.A., E.C., W.K., K.-K.N., and J.M.D. contributed to the experimental effort. All authors discussed the results and contributed to the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: All data needed to evaluate the conclusions in this paper are present in the paper or in the supplementary materials. All data presented in this paper are deposited at Zenodo (65). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf8999 Supplementary Text Figs. S1 to S3 Reference (66) Submitted 18 November 2022; accepted 12 October 2023 10.1126/science.adf8999 Bao et al., Science 382, 1138–1143 (2023) 8 December 2023 6 of 6
10.1126_science.adi2436
RES EARCH MOLECULAR BIOLOGY Human POT1 protects the telomeric ds-ss DNA junction by capping the 5′ end of the chromosome Valerie M. Tesmer, Kirsten A. Brenner, Jayakrishnan Nandakumar* Protection of telomeres 1 (POT1) is the 3′ single-stranded overhang-binding telomeric protein that prevents an ataxia telangiectasia and Rad3–related (ATR) DNA damage response (DDR) at chromosome ends. What precludes the DDR machinery from accessing the telomeric double-stranded–single-stranded junction is unknown. We demonstrate that human POT1 binds this junction by recognizing the phosphorylated 5′ end of the chromosome. High-resolution crystallographic structures reveal that the junction is capped by POT1 through a “POT-hole” surface, the mutation of which compromises junction protection in vitro and telomeric 5′-end definition and DDR suppression in human cells. Whereas both mouse POT1 paralogs bind the single-stranded overhang, POT1a, not POT1b, contains a POT-hole and binds the junction, which explains POT1a’s sufficiency for end protection. Our study shifts the paradigm for DDR suppression at telomeres by highlighting the importance of protecting the double-stranded–single-stranded junction. N ucleoprotein complexes called telomeres cap chromosome ends to ensure genome integrity. Human telomeric DNA contains ~10 to 15 kb of tandem 5′-GGTTAG-3′/3′- CCAATC-5′ repeats. Although telomeric DNA is primarily double-stranded (ds), all chromosomes terminate in a 50- to 500- nucleotide (nt) single-stranded (ss) G-rich telomeric overhang (Fig. 1A, bottom) (1). The six-protein shelterin complex coats telomeric DNA to protect chromosome ends from being recognized as dsDNA breaks by the ataxia telangiectasia and Rad3–related (ATR) kinase– and ataxia-telangiectasia mutated (ATM) kinase–mediated DNA damage response (DDR) machineries (2, 3). ATR signaling involves multiple protein factors and coordinated recog- nition of both the ss and the adjacent ds-ss junction of its DNA substrates (4). Protection of telomeres 1 (POT1) is a shelterin component that binds the ss G-rich overhang with high affinity and sequence specificity and prevents ATR signaling at telomeres (2, 5, 6). POT1 recognizes ssDNA through its DNA bind- ing domain (DBD), which consists of two oligonucleotide/oligosaccharide-binding (OB) domains (Fig. 1A). Previous studies have re- ported a decanucleotide TTAGGGTTAG within two telomeric ss repeats, 1GGTTAGGGTTAG12, to be sufficient for high-affinity binding to hu- man POT1 (hPOT1) (7). The first OB domain (OB1) of hPOT1 binds 3TTAGGG8 (OB1DNA), whereas its second OB domain (OB2) binds 9TTAG12 (OB2DNA) (Fig. 1, A and B) (7). Homologs of POT1 are identifiable across eukaryotes (5, 8–15), and deleting the POT1 paralog in mice that is involved in chromosome-end pro- tection (POT1a) is embryonic lethal (14, 15). Department of Molecular, Cellular and Developmental Biology, University of Michigan, Ann Arbor, MI 48109, USA. *Corresponding author. Email: jknanda@umich.edu The current model for ATR suppression at telomeres invokes the prevention by POT1 of replication protein A (RPA) loading onto the ss overhang, through POT1’s high affinity for telomeric ssDNA and its tethering to the rest of shelterin at telomeric dsDNA (2, 6, 16). Yet, multiple observations suggest that addition- al features of POT1 are involved in ATR rep- ression. First, mouse POT1 paralogs POT1a and POT1b display indistinguishable ssDNA- binding activity, but only POT1a is sufficient for chromosome-end protection (6, 15, 17), whereas POT1b regulates chromosome-end processing and replication activities (16, 18–21). Replac- ing the DBD of POT1b with that of POT1a or hPOT1 enables ATR repression at telomeres (17). Second, replacing the DBD of POT1a with that of ssDNA-binding protein RPA70 is not sufficient to fully repress ATR signaling at telomeres in mouse cells that lack POT1a (16). Moreover, POT1’s binding to the G-rich ss overhang does not explain how it dictates the 5′ end of the C-rich strand, which terminates predominantly in ATC-5′ in mammals (22, 23) (Fig. 1A, bottom). These observations are con- sistent with the DBD of hPOT1 and mouse POT1a carrying out an additional function relevant to ATR suppression. Results Human POT1 binds a 5′-phosphorylated telomeric ds-ss DNA junction We hypothesized that hPOT1 binds to the telo- meric ds-ss junction after we reanalyzed its pub- lished DNA binding-site preferences. Two classes of POT1 binding sites emerged from previous SELEX (systematic evolution of ligands by ex- ponential enrichment) analysis, one of which was the expected 3TTAGGGTTAG12 (OB1DNA and OB2DNA) site (Fig. 1B, Class I) (24). A second class contained OB1DNA, an upstream tri-K (“K” indi- cates a G or T nucleotide), and a seemingly non- telomeric (NT) sequence implicated in binding to OB1 (consensus: CTCCAGCAGGGG3TTAGGG8) (Fig. 1B, Class II) (24). Junction binding was suspected on the basis of the observation that the tri-K GGG motif corresponds to the telo- meric repeat sequence upstream of OB1DNA, and NT sequences in the Class II hits could fold into a hairpin (hp) containing a 2-base-pair (bp) stem −1GG0/−6CC−7 and a variable tetraloop (positions −5 to −2) (Fig. 1B and fig. S1, A and B). In this interpretation, G0 and C−7 represent the first base pair at the ds-ss junction (with C−7 corresponding to the 5′ end of the mammalian chromosome), and the 3′ overhang initiates in the GGTTAG register (Fig. 1, A and B, and fig. S1, A and B). We conducted a quantitative electro- phoretic mobility shift assay (EMSA) with puri- fied hPOT1 DBD (hDBD) (fig. S2A) and a 5′-32P– labeled hp oligonucleotide derived from the Class II consensus terminating in a C at the 5′ end and containing a 3′ overhang of sequence 1GGTTAGGG8 (hp-ss1-8) (Fig. 1C). The absence of OB2DNA from the Class II consensus atten- uates the affinity of hDBD for ssDNA (7), allow- ing us to assess DNA affinity of POT1 for the ds-ss junction. hDBD bound strongly to hp-ss1-8 [dissociation constant (Kd) = 2.6 ± 0.3 nanomolar (nM)] but not to a similar target (no_hp-ss1-8) that lacks the ability to form a hairpin (Fig. 1, C and D). The natural telomeric ds-ss junction ends in a 5′-phosphate (5′-P), which has been previously exploited to determine the 5′-terminal nt of chromosomes by using DNA ligase- mediated methods (25). To test the importance of this phosphate in binding hPOT1, we per- formed a competition experiment mixing 5′-32P-hp-ss1-8 with either nonradiolabeled 5′-phosphorylated hp-ss1-8 or 5′-OH-hp-ss1-8 be- fore binding to hDBD. The 5′-P was required to effectively outcompete POT1 binding to the radiolabeled DNA (Fig. 1E). The absence of a 5′-P at the DNA junction in past in vitro studies may have prevented the detection of this previously unappreciated POT1 DNA- binding activity (17, 26–29). POT1 bound to a telomeric ds-ss junction in vivo is poised to engage both OB1DNA and OB2DNA. Extension of the overhang of the hp to include OB2DNA (hp- ss1-12) resulted in a higher affinity for hDBD [Kd = 70 picomolar (pM)] (Fig. 1, C and F) compared with either ss1-12 (Kd = 190 pM) (Fig. 1C and fig. S2B) or hp-ss1-8 (Fig. 1, C and D). We confirmed that a heterodimer of full-length hPOT1 and shelterin partner TINT1-PTOP-PIP1 (TPP1) (by using the TPP1N truncation con- struct), which approximates the context of hPOT1 coating the ss overhang in vivo (30, 31), exhibited robust binding to 5′-P-hp-ss1-12 (Fig. 1G). DBD bound a two-stranded DNA (duplex reinforced with 30 bp of arbitrary, nontelomeric sequence) terminating in 5 bp of native ds telomeric junction sequence and an 8-nt overhang (long_ds-ss1-8) with an affinity that was approximately one order of magni- tude greater than observed with hp-ss1-8, likely Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Human POT1 recognizes the 5′-phosphorylated ds-ss junction of telomeres. (A) (Top) Schematic of hPOT1 includes binding domains for ssDNA (hDBD) and TPP1 (TPP1-BD). hDBD (PDB: 1XJV) is composed of OB1 and OB2. The current model suggests that POT1 outcompetes the ssDNA-binding RPA complex to prevent ATR signaling at telomeres. HJRL, Holliday junction resolvase-like. (Bottom) Mammalian chromosomes end in a ds-ss junction containing ATC-5′ (predominantly) and a ss G-rich overhang. Numbering starts with the first overhang nucleotide. (B) A previous SELEX study revealed two hPOT1-binding DNA classes (24). Class I harbors the known sites for OB1 and OB2, denoted as OB1DNA (cyan) and OB2DNA (pink), respectively. Class II revealed a consensus containing a seemingly nontelomeric (NT) sequence upstream of OB1DNA that can potentially fold into a hp; “K” indicates a G or T nucleotide, and the shaded area indicates the sequence of the first bp at the telomeric ds-ss junction. (C) Annotated name, sequence, predicted hp structure [with Tm (where Tm is the temperature at which 50% of dsDNA is denatured) calculated by the UNAFold web server], and mean Kd and SD (of binding to hDBD) of the oligonucleotides used in EMSA analysis. NA indicates not applicable. (D to H) EMSA of indicated proteins (hDBD or POT1- TPP1N heterodimer) and 5′-32P–labeled DNA oligo- nucleotides. (D), (F), (G), and (H) indicate direct binding experiments, and (E) indicates a competition experiment. In (D), 0.1 nM 5′-32P-hp-ss1-8 was used; the number of experimental replicates n = 5 for hp-ss1-8 (full and partial titrations); n = 3 for no_hp-ss1-8. In (E), 100 nM hDBD and 0.1 nM 5′-32P–labeled hp-ss1-8 were incubated with indicated amounts of unlabeled hp-ss1-8 (cold DNA) containing either a 5′-OH or a 5′-P; n = 3. In (F), 0.001 nM 5′-32P-hp-ss1-12 was used; n = 3. In (G), 0.01 nM 5′-32P-hp-ss1-12 was used; n = 3. (H) 0.001 nM 5′-32P-long_ds-ss1-8 was used; n = 3. Circled red “P” indicates radiolabeled; circled black “P” indicates nonradiolabeled. Bound (B) indicates DNA bound to protein; Free (F) indicates free, unbound DNA. reflecting the greater stability of the more phys- iologically representative duplex DNA versus that of the hp (Fig. 1, C and H). Our data dem- onstrate that the telomeric ds-ss junction is a previously unappreciated high-affinity binding site for hPOT1. High-resolution structures reveal how human POT1 caps the phosphorylated 5′ end of a telomeric junction To determine the structural basis for hPOT1’s telomeric ds-ss junction–binding activity, we formed complexes of hDBD with two sub- strates that mimic the telomeric ds-ss junction— 5′-P-ds-ss1-12 (DNA containing a 5-bp arbitrary, nontelomeric tether upstream of GTTAG/CAATC- 5′-P native telomeric ds sequence extending into a 12-nt 3′ overhang) (fig. S2, C and D) and 5′-P-hp- ss1-12 (Fig. 1C)—and solved their structures using x-ray crystallography (Fig. 2, A and B). The hDBD-bound 5′-P-ds-ss1-12 and 5′-P-hp-ss1-12 structures were solved to 2.60- and 2.16-Å res- olution, respectively (table S1). Both struc- tures are similar to each other (fig. S3D) and recapitulate the previously reported hDBD-ss DNA-binding interface with minor differences (fig. S3, A to C, and E to J) (7). These structures reveal how hPOT1 binds the phosphorylated 5′ end of the telomeric ds-ss junction (Fig. 2). An electropositive pocket of four amino acids (Y9, R80, H82, and R83) in the hPOT1 OB1 domain that we name the “POT-hole” caps the 5′-P- cytidine nucleotide by means of a network of stacking and electrostatic interactions (Fig. 2, D to F, and fig. S4A). R83 acts as the linchpin by forming an ionic interaction with the 5′-P, Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Structural basis of telomeric junction 5′- end protection by human POT1. (A and B) Cartoon representation of high-resolution crystal structures of complexes of (A) hDBD with 5′-P-ds- ss1-12 and (B) 5′-P-hp-ss1-12, showing OB1 (cyan) and OB2 (pink) bound to DNA [gray, with the exception of the 5′-P, whose atoms are shown as spheres and in Corey-Pauling-Koltun (CPK) coloring]. A boxed schematic of the DNA is shown below the structure, with the ds sequence found naturally at the telomeric ds-ss junction shaded gray, the 5′-P in red, and residues in the G-rich 3′ overhang colored to indicate binding by OB1 and OB2, respectively. (C) Cartoon and (E) electrostatic surface (blue is electropositive and red is electronegative) repre- sentations of the hDBD-5′-P-ds-ss1-12 structure shown in a view orthogonal to that in (A). The 5′-P occupies a pocket in POT1 that is complementary in shape and charge. Single-letter abbreviations for the amino acid residues are as follows: H, His; R, Arg; and Y, Tyr. (D) The POT-hole-DNA interface within the hDBD-5′-P-hp-ss1-12 structure is shown with POT-hole side chains (carbon in cyan) and the nucleotides (carbon in light gray) near the junction shown as sticks. A water molecule bridging hPOT1 R80 to the 5′-P is shown as an orange sphere. The dashed lines indicate H-bonds and ionic interactions, the double-headed arrow indicates stacking of the hPOT1 R83 side chain with the 5′-C at the junction (numbered C0), and N indicates the N terminus of hDBD resolved in the crystal structure. (F) Interaction map of hDBD with the ds-ss junction. stacking against the 5′-cytosine base, and form- ing hydrogen bonds (H-bonds) with the ribose- ring oxygen of the 5′-cytidine nucleoside (5′-C) (Fig. 2, D and F, and fig. S4A). R83 also forms an H-bond with Y9, which along with H82 interacts with the 5′-P. R80 forms a water- mediated H-bond with the 5′-P. We observed that the POT-hole is not optimally sized to accommodate a bulkier adenine (purine in- stead of a pyrimidine) or thymine (methyl group on the base) at the 5′ end because of steric clashes (fig. S5, A to C). Furthermore, the fixed distance between the POT-hole and the ssDNA-binding region of hPOT1 dictates the preference for the naturally occurring ATC-5′ versus alternative 5′-C iterations: ATCC-5′ and ATCCC-5′ (fig. S5D). In addition to interactions involving the POT- hole, junction recognition is fortified by con- tacts made by the backbone amides of hPOT1 amino acids 121 to 124 with the phosphodiester group penultimate to the 5′-C (Fig. 2F and fig. S4B), as well as S99 with G2 (Fig. 2F and fig. S3, K and L). These data provide the structural basis for binding of the telomeric ds-ss junction by hPOT1. The POT-hole dictates telomeric DNA junction binding and inhibits DDR at telomeres We evaluated the importance of the POT-hole in binding the telomeric ds-ss junction in vitro using purified hDBD variants with alanine mutations at Y9, R80, H82, and R83 (fig. S6A). We also engineered an R83E charge-reversal mutant to test the importance of the ionic R83- 5′-P interaction. Alanine substitution of F62, a residue in hPOT1 OB1 that is indispensable for binding telomeric ssDNA (32), was included as a control to disrupt binding to both ssDNA and the ds-ss junction. In agreement with the structural data, little to no DNA binding was observed for any POT-hole mutant with the 5′-P-hp-ss1-8, even at concentrations 100-fold greater than the Kd with wild-type (WT) hDBD (Fig. 3A, left). By contrast, POT-hole mutants bound 5′-P-ss1-12 with an affinity similar to that of wild type (Fig. 3A, right, and fig S6, B and C). F62A failed to bind either oligonucleotide, which is consistent with binding to OB1DNA Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Separation-of-function POT-hole mutations abrogate ds-ss DNA junction binding in vitro and result in a DDR at human telomeres. (A) EMSA to detect direct binding of WT or indicated mutant hDBD constructs with 5′-32P-hp-ss1-8 (0.1 nM; lanes 1 to 22) and 5′-32P-ss1-12 (0.1 nM; lanes 23 to 30); n = 3. (B) Schematic conveying how POT-hole mutations would disrupt binding to the ds-ss junction but not coating of the ss overhang by POT1. (C) Scheme for deletion of endogenous POT1 and complementation with lentivirally transduced hPOT1-Myc to assess the ability of mutants to suppress TIF formation in a HEK 293E–based cell line (34). (D) TIF analysis of cell lines after 4-OHT and dox (1000 ng/ml; 25 ng/ml in “low dox” wild type) treatment as described in (C) performed with peptide nucleic acid fluorescent in situ hybridization (PNA-FISH) for telomeres (green) and immunofluorescence (IF) for Myc (hPOT1; cyan) and 53BP1 (red). 4′,6- diamidino-2-phenylindole (DAPI) was used to stain the nucleus (blue). Overlap of the telomeric and 53BP1 foci (and Myc foci, if applicable) in the “Merge” panel indicates TIFs. (Inset) Magnified view of the boxed area within the image; arrowheads indicate TIFs. (E) Quantitation of TIF data of which (D) is representative. Mean and SD (n = 3 for all conditions except WT −dox, for which n = 5; each +dox set containing >75 nuclei and each −dox set containing >50 nuclei) for TIFs are plotted for the indicated cell lines. P values calculated with a two-tailed Student’s t test for comparisons against WT +dox data are indicated above the bars. Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Presence of the POT-hole dictates POT1 paralog choice for chromosome-end protection in mice. (A) Human POT-hole residues are conserved in mouse POT1a but not mouse POT1b. (B) Electrostatic surface comparisons of hDBD (from hDBD-5′-P-ds-ss1-12 structure) and POT1a and POT1b DBD (Alphafold models), with the phosphorylated 5′-C of the hDBD-bound structure shown in sticks. (C and D) EMSA analysis of indicated mouse POT1a and POT1b DBD constructs with the indicated 5′-32P–labeled oligonucleotides [0.1 nM for (C) and (D), right; 0.01 nM for (D), left]; n = 3. (E) EMSA analysis of indicated human and mouse POT1 DBD constructs with 0.001 nM 5′-32P-long_ds-ss1-8 two-stranded DNA; n = 3. (F) (Top left) Names and sequences of the two DNA oligonucleotides, hp-ss1-24 and ss1-24, used to evaluate 5′-end–binding prefer- ence. Both DNAs were labeled at the 5′ end with 32P for EMSA analysis. (Bottom left) Three possible DNA-binding registers for the first DBD molecule are shown with the center-binding register precluding the binding of a second DBD molecule. (Right) EMSA analysis of POT1a DBD with hp-ss1-24 (discrete slow-migrating band with increasing concentrations of protein; 2×B) and ss1-24 (smeary band; mixture of B and 2×B), DNA at 0.1 nM; n = 3. (G) EMSA analysis of indicated POT1a DBD constructs with 0.1 nM hp-ss1-24. YHR, triple mutant Y9S-H82Q-R83G; n = 3. being critical for both DNA-binding modes. These data highlight the importance of the POT- hole in 5′-end binding and provide separation- of-function mutants to test the importance of hPOT1’s junction-binding activity in cells (Fig. 3B). Loss of POT1 binding at the 3′ overhang re- sults in telomere dysfunction–induced foci (TIF), which signify the recognition of telomeres by the DDR machinery (33). To determine the bio- logical importance of the POT-hole binding to the telomeric junction, we used a previously described cell line in which POT1 can be deleted in an inducible fashion (POT1 KO) (34) to test the ability of transduced WT and mutant hPOT1 Myc-tagged constructs to compensate for the loss of endogenous POT1 (materials and meth- ods). Transduced cells were treated first with 4-OHT to delete POT1 and then either treated with doxycycline (dox) to induce exogenous hPOT1 expression (“+dox”) or left untreated (“−dox”) (Fig. 3C). In the absence of dox, 4-OHT treatment resulted in a robust TIF phenotype, characterized by colocalization of the DDR factor 53BP1 at telomeres (fig. S6, E and F). hPOT1 wild type and “low dox” wild type, but not hPOT1 F62A, suppressed TIFs (Fig. 3, D and E). POT-hole mutants Y9A, R83A, and R83E were defective in TIF sup- pression compared with wild type, with R83E being the most deleterious (Fig. 3, D and E). This trend emphasizes the importance of the ionic interaction between R83 and the 5′-P. Clones isolated from 6X-Myc–tagged hPOT1 WT, F62A, and R83E cell populations also reca- pitulated the TIF phenotypes (fig. S7, A to D). Furthermore, TIFs were smaller (Fig. 3D, inset) and less frequent (Fig. 3, D and E) in POT-hole mutant cells compared with those in F62A Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. Maintenance of the ATC-5′ end of chromo- somes by the POT-hole. (A) Schematic of the modified STELA technique for determining the chromosomal 5′-terminal nucleotide in human cell lines. Step 1: DNA ligation of genomic DNA 5′-P ends with telorettes ending in each of the six possible repeat registers at the 3′-OH ends. Step 2: PCR amplification of the ligation products performed with a forward primer (PCR-F) targeting the subtelomere of chromosome XpYp and a reverse primer (PCR-R) targeting a sequence shared by all telorettes. The products are visualized with Southern blot analysis performed with a 5′-32P–labeled XpYpB2 reverse primer. (B) STELA-based determination of the chromosomal 5′-terminal nucleotide in the HEK 293E–based POT1 KO parental cell line (−4-OHT and +4-OHT) and hPOT1-Myc WT– or R83E- complemented clonal cell lines treated with both 4-OHT and dox. (C) Quantitation of ATC-5′ preference calculated as the ratio of the total band intensity in the primer 3 lane over the total intensity over all six lanes. Mean and SD for n = 4 replicates of which B is representative are plotted. P values were calculated with a two-tailed Student’s t test for comparisons against parental −4-OHT data (for parental +4-OHT) or hPOT1-Myc WT clones (for hPOT1-Myc R83E clones). (D) (Left) TRF analysis of cell lines used in (B) performed first under native conditions with a 5′-32P–labeled telomeric C-probe (CTAACC)4 to detect the ss G-rich overhang. (Right) TRF analysis after denaturing the DNA on the same gel and reprobing it to detect the total telomeric DNA signal; n = 1. (E) Model for ATR inhibition at telomeres by POT1. The ssDNA- binding of hPOT1 prevents the loading of RPA to curb ATR recruitment to the 3′ overhang. Protection of the ds-ss junction by hPOT1 prevents loading of the 9-1-1/Rad17-RFC clamp and clamp-loader complex and ATR activator TOPBP1. In mice, both POT1 paralogs coat the ss overhang, but only POT1a protects the ds-ss junction. The shelterin proteins protecting the telomeric dsDNA are expected to keep POT1-TPP1 tethered to the ss overhang, facilitated by protein-protein interactions and the conformational flexibility within the proteins (29) and the telomeric DNA. cells. This finding suggests that both junction- and ssDNA-binding activities of hPOT1 must be compromised to trigger a full DDR (see Discussion). Our results demonstrate that junc- tion binding, which should involve a single POT1 molecule per chromosome end (Fig. 3B), is critical for chromosome-end protection. The POT-hole differentiates mouse POT1 paralogs and enables POT1a to protect the telomeric junction Despite being strictly conserved in other mam- malian POT1 homologs, including mouse POT1a, each of the four POT-hole amino acids is re- placed with a structurally disparate residue in mouse POT1b (Fig. 4A and fig. S8A). By contrast, the residues used in ss DNA binding are con- served in all mammalian POT1 homologs, in- cluding POT1b (fig. S8A). Aligning Alphafold predictions (35) of POT1a and POT1b DBDs with the junction-bound structure of hDBD illustrates that the shape and electropositive nature of the POT-hole are predicted to be lost in POT1b (Fig. 4B and fig. S8, B and C). We hypothesized that POT1a, but not POT1b, pro- tects the 5′ end at the junction. Indeed, POT1a- and POT1b-DBD proteins bound ss1-12, but only POT1a DBD engaged a telomeric ds-ss junction with high affinity (Fig. 4, C to E, and fig. S8, D and E). POT1a replaced with POT1b residues in the POT-hole (except R80; fig. S8F legend explains rationale) retained affinity toward ss1-12 (Fig. 4C) but failed to bind the junction (Fig. 4D, right, and E, and fig. S8D). To measure junction-binding in the presence of multiple ss DNA-binding sites, we developed an EMSA-based “POT1 packing” assay with two DNA targets, each containing four telomeric ss repeats (24 nt) spanning three possible Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E POT1-binding registers. The 5′ and 3′ registers are compatible with the packing of two POT1 molecules, whereas binding to a central reg- ister precludes the loading of a second POT1 (Fig. 4F, left). hp-ss1-24 includes a ds-ss junction upstream of this ss region, whereas ss1-24 does not. A fully packed 2:1 DBD-DNA complex would produce a sharp, slow-migrating band at higher DBD concentrations, whereas a mix of 2:1 and 1:1 complexes (of various binding registers) would generate a smear. POT1a DBD binding resulted in a sharp band for hp-ss1-24 but not ss1-24, suggesting that the protein packs preferentially against a ds-ss junction but that there is no end-binding bias to dissuade it from binding to the central site of ss1-24 (Fig. 4F, right). POT1a POT-hole mutants R83G and triple mutant YHR lost the ability to pack at the junction (Fig. 4G), which is consistent with R83 capping the 5′ terminus (Fig. 2, D and F) and repressing TIFs (Fig. 3, D and E). hDBD and mouse POT1b DBD formed a discrete complex with not only hp-ss1-24 but also ss1-24, which is consistent with a 3′-end–binding preference (fig. S9, A and B) (7). Our results demonstrate that the POT-hole allows POT1a to preferen- tially bind the telomeric junction. The POT-hole helps maintain the 5′-end identity of human chromosomes Consistent with the structures we solved, the POT-hole of hDBD and mouse POT1a DBD pro- tect the 5′-P end from 5′ exonucleolytic action in vitro (fig. S10, A to F). We next asked whether the POT-hole helps maintain the 5′-terminal sequence of the chromosomes in cells. We used a modified single telomere length analysis (STELA) approach that uses ligation-mediated polymerase chain reaction (PCR) amplification to determine the abundance of each of the six possible chromosomal 5′-end permutations (Fig. 5A) (23). Genomic DNA extracted from the parental human embryonic kidney (HEK) 293E cell line displayed the expected ATC-5′ prefer- ence that is lost after POT1 deletion (Fig. 5, B and C). WT hPOT1, but not R83E hPOT1, was able to restore the ATC-5′ bias to untreated (parental –4-OHT) levels, demonstrating that the POT-hole helps maintain the 5′ end of the human chromosome (Fig. 5, B and C, and fig. S10G). The 5′-end scrambling of hPOT1 R83E was less severe than that of POT1 KO. This dif- ference may be explained by the unleashing of 5′ exonuclease activity at telomeres completely devoid of POT1 (36). Terminal restriction frag- ment (TRF) analysis reproduced previously characterized phenotypes (15, 34, 37), including the accumulation of slow-migrating species (denatured and native blots) and an increase in the G-rich ss signal (native blot) upon POT1 deletion, which were suppressed by expression of hPOT1 wild type but not F62A (Fig. 5D and fig. S10H). R83E recapitulated the WT pheno- types, suggesting that the end-protection func- tion of the POT-hole is separable from hPOT1’s role in bulk-telomere or overhang-length main- tenance. Thus, the POT-hole helps maintain ATC-5′ ends without acutely influencing telo- mere length. Discussion The major pathway of ATR activation requires RPA binding to exposed ssDNA and recogni- tion of the ds-ss junction by the 9-1-1/Rad17-RFC (RAD9–RAD1–HUS1/Rad17-RFC2–RFC3–RFC4– RFC5) clamp and clamp loader, which with the MRN (MRE11-RAD50-NBS1) complex recruit TOPBP1 (DNA topoisomerase 2-binding pro- tein 1) to activate ATR (Fig. 5E) (4, 38). The struc- ture of human 9-1-1/Rad17-RFC bound to a ds-ss junction revealed a basic pocket in Rad17 that is poised to bind the 5′-phosphorylated end of a junction by using a mechanism similar to that of POT1 (fig. S11, A and B) (39). Con- sistent with a competition between POT1 and 9-1-1/Rad17-RFC in binding the ds-ss junction, subunits of the 9-1-1 and MRN com- plexes, as well as TOPBP1, are enriched at telomeres in the absence of hPOT1 (34). We therefore propose that POT1 not only out- competes RPA at the telomeric ss overhang but also prevents ATR activation by denying 9-1-1/Rad17-RFC access to the telomeric ds-ss junction (Fig. 5E). The duplication of POT1 (40), the conserva- tion of the POT-hole in POT1a (fig. S12A), the disruption of the POT1-hole in POT1b (fig. S12B), and the retention of CTC1-STN1-TEN1 (CST)– binding motifs in POT1b (40) within the Muri- dae and Cricetidae families of the Rodentia order provide support to the hypothesis that POT1b relinquished junction binding to facili- tate processes at the 3′ end. We propose that POT1a wards off 9-1-1/Rad17-RFC at the junc- tion, although both POT1a and POT1b paralogs could counter RPA at the overhang in mouse cells (Fig. 5E). The POT-hole is conserved in species dis- tant to mammals, such as Sterkiella nova and Caenorhabditis elegans (fig. S13A). The precisely defined S. nova macronuclear telomere contains a 5′-C at the ds-ss junction and a 16-nt overhang that binds one telomere end–binding protein (TEBP)a/b complex (homologous to the POT1- TPP1 complex) (41). TEBPa has been crystal- lized with a sulfate ion bound in a manner indistinguishable from how the 5′-P binds hDBD in our junction-bound structures (fig. S13B) (42). Indeed, like hPOT1, TEBPa binds the telomeric ds-ss junction more strongly than it binds telo- meric ssDNA (8). These observations point to a single TEBPa/b complex simultaneously pro- tecting the 5′ and 3′ ends of the chromosome (8, 41, 42). Schizosaccharomyces pombe, in which a POT-hole is not obvious (fig. S13, A and C) (5, 43), and eukaryotes whose chromosomes do not end in a 5′-C, must have evolved alternative approaches for junction protection. We updated the model for how telomeres avert detection by the DDR machinery to include a critical role of POT1 in binding the telomeric ds-ss junction. Thus, POT1 protects both DNA strands at human chromosome ends by coating the G-rich ss overhang and recognizing the phos- phorylated 5′ end of the C-rich strand. REFERENCES AND NOTES 1. W. Palm, T. de Lange, Annu. Rev. Genet. 42, 301–334 (2008). 2. E. L. Denchi, T. de Lange, Nature 448, 1068–1071 (2007). 3. T. de Lange, Annu. Rev. Genet. 52, 223–247 (2018). 4. J. C. Saldivar, D. Cortez, K. A. Cimprich, Nat. Rev. Mol. Cell Biol. 18, 622–636 (2017). 5. P. Baumann, T. R. Cech, Science 292, 1171–1175 (2001). 6. Y. Gong, T. de Lange, Mol. Cell 40, 377–387 (2010). 7. M. Lei, E. R. Podell, T. R. Cech, Nat. Struct. Mol. Biol. 11, 1223–1229 (2004). 8. D. E. Gottschling, V. A. Zakian, Cell 47, 195–205 (1986). 9. C. M. Price, T. R. Cech, Genes Dev. 1, 783–793 (1987). 10. M. P. Horvath, V. L. Schweiker, J. M. Bevilacqua, J. A. Ruggles, S. C. Schultz, Cell 95, 963–974 (1998). 11. C. I. Nugent, T. R. Hughes, N. F. Lue, V. Lundblad, Science 274, 249–252 (1996). 12. J. J. Lin, V. A. Zakian, Proc. Natl. Acad. Sci. U.S.A. 93, 13760–13765 (1996). 13. M. Raices et al., Cell 132, 745–757 (2008). 14. L. Wu et al., Cell 126, 49–62 (2006). 15. D. Hockemeyer, J. P. Daniels, H. Takai, T. de Lange, Cell 126, 63–77 (2006). 16. K. Kratz, T. de Lange, J. Biol. Chem. 293, 14384–14392 (2018). 17. W. Palm, D. Hockemeyer, T. Kibe, T. de Lange, Mol. Cell. Biol. 29, 471–482 (2009). 18. P. Gu et al., Nat. Commun. 12, 5514 (2021). 19. L. Y. Chen, S. Redon, J. Lingner, Nature 488, 540–544 (2012). 20. P. Wu, H. Takai, T. de Lange, Cell 150, 39–52 (2012). 21. M. Wan, J. Qin, Z. Songyang, D. Liu, J. Biol. Chem. 284, 26725–26731 (2009). 22. D. Hockemeyer, A. J. Sfeir, J. W. Shay, W. E. Wright, T. de Lange, EMBO J. 24, 2667–2678 (2005). 23. A. J. Sfeir, W. Chai, J. W. Shay, W. E. Wright, Mol. Cell 18, 131–138 (2005). 24. K. H. Choi et al., Biochimie 115, 17–27 (2015). 25. D. M. Baird, J. Rowson, D. Wynford-Thomas, D. Kipling, Nat. Genet. 33, 203–207 (2003). 26. C. J. Lim, A. J. Zaug, H. J. Kim, T. R. Cech, Nat. Commun. 8, 1075 (2017). 27. T. Paul, W. Liou, X. Cai, P. L. Opresko, S. Myong, Nucleic Acids Res. 49, 12377–12393 (2021). 28. S. Ray, J. N. Bandaria, M. H. Qureshi, A. Yildiz, H. Balci, Proc. Natl. Acad. Sci. U.S.A. 111, 2990–2995 (2014). 29. J. C. Zinder et al., Proc. Natl. Acad. Sci. U.S.A. 119, e2201662119 (2022). 30. F. Wang et al., Nature 445, 506–510 (2007). 31. J. Nandakumar et al., Nature 492, 285–289 (2012). 32. H. He et al., EMBO J. 25, 5180–5190 (2006). 33. H. Takai, A. Smogorzewska, T. de Lange, Curr. Biol. 13, 1549–1556 (2003). 34. G. Glousker, A. S. Briod, M. Quadroni, J. Lingner, EMBO J. 39, e104500 (2020). 35. J. Jumper et al., Nature 596, 583–589 (2021). 36. T. Kibe, M. Zimmermann, T. de Lange, Mol. Cell 61, 236–246 (2016). 37. D. Loayza, T. De Lange, Nature 423, 1013–1018 (2003). 38. C. A. MacDougall, T. S. Byun, C. Van, M. C. Yee, K. A. Cimprich, Genes Dev. 21, 898–903 (2007). 39. M. Day, A. W. Oliver, L. H. Pearl, Nucleic Acids Res. 50, 8279–8289 (2022). 40. L. R. Myler et al., Genes Dev. 35, 1625–1641 (2021). 41. L. A. Klobutcher, M. T. Swanton, P. Donini, D. M. Prescott, Proc. Natl. Acad. Sci. U.S.A. 78, 3015–3019 (1981). 42. S. Classen, J. A. Ruggles, S. C. Schultz, J. Mol. Biol. 314, 1113–1125 (2001). 43. M. Lei, E. R. Podell, P. Baumann, T. R. Cech, Nature 426, 198–203 (2003). AC KNOWLED GME NTS We thank S. Padmanaban (Nandakumar laboratory) for input on the design of the cell-based experiments; G. Glousker and J. Lingner [Ecole Polytechnique Fédérale de Lausanne (EPFL), Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E Switzerland] for graciously gifting the POT1-inducible KO HEK 293E cell line and the pCW22_TREtight_MCS_UBC_rtTA_IRES_Blast lentiviral vector for dox-inducible expression and for sharing detailed protocols and troubleshooting tips for these resources; J. Schmidt (Michigan State University, USA) for the 53BP1 antibody; the beamline staff at the Life Sciences Collaborative Access Team (LS-CAT) beamline of the Argonne National Laboratory for help with x-ray diffraction data collection [use of the Advanced Photon Source, an Office of Science User Facility operated for the US Department of Energy (DOE) Office of Science by Argonne National Laboratory, was supported by the US DOE under contract no. DE-AC02-06CH11357]; F. C. Lowder and L. Simmons (University of Michigan at Ann Arbor, USA) for helpful suggestions for the exonuclease protection experiment carried out with fluorophore- labeled DNA and for preparation of an RNA ladder; T. de Lange and S. Cai (Rockefeller University, USA), J. Schmidt (Michigan State University, USA), H. Shibuya (University of Gothenburg, Sweden), and J. Williams (Nandakumar laboratory) for helpful comments on the manuscript; and G. Sobocinski for help with microscopy. Funding: This study was supported by NIH grants R01GM120094, R01HD108809, and R35GM148276 (J.N.) and by American Cancer Society Research Scholar grant RSG-17-037-01-DMC (J.N.). Author contributions: Conceptualization: V.M.T. and J.N. Methodology: V.M.T., K.A.B., and J.N. Investigation: V.M.T. and K.A.B. Visualization: V.M.T., K.A.B., and J.N. Funding acquisition: J.N. Project administration: V.M.T. and J.N. Supervision: V.M.T. and J.N. Writing – original draft: V.M.T. and J.N. Writing – review and editing: V.M.T., K.A.B., and J.N. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data are available in the main and supplementary figures. All material generated in this study, such as plasmids for protein expression and cell lines, are available upon request. Coordinates and structure factors of the crystal structures of hDBD with 5′-P-hp-ss1-12 and 5′-P- ds-ss1-12 are deposited in the Protein Data Bank (PDB) under accession codes 8SH0 and 8SH1, respectively. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi2436 Materials and Methods Figs. S1 to S13 Tables S1 and S2 References (44–55) MDAR Reproducibility Checklist Submitted 19 April 2023; accepted 19 July 2023 10.1126/science.adi2436 Tesmer et al., Science 381, 771–778 (2023) 18 August 2023 8 of 8
10.1126_science.adh9059
RES EARCH HOST-GUEST CHEMISTRY Disequilibrating azobenzenes by visible-light sensitization under confinement Julius Gemen1, Jonathan R. Church2, Tero-Petri Ruoko3, Nikita Durandin3, Michał J. Białek4, Maren Weißenfels1, Moran Feller1, Miri Kazes1, Magdalena Odaybat5, Veniamin A. Borin2, Rishir Kalepu1, Yael Diskin-Posner6, Dan Oron1, Matthew J. Fuchter5, Arri Priimagi3, Igor Schapiro2*, Rafal Klajn1,7* Photoisomerization of azobenzenes from their stable E isomer to the metastable Z state is the basis of numerous applications of these molecules. However, this reaction typically requires ultraviolet light, which limits applicability. In this study, we introduce disequilibration by sensitization under confinement (DESC), a supramolecular approach to induce the E-to-Z isomerization by using light of a desired color, including red. DESC relies on a combination of a macrocyclic host and a photosensitizer, which act together to selectively bind and sensitize E-azobenzenes for isomerization. The Z isomer lacks strong affinity for and is expelled from the host, which can then convert additional E-azobenzenes to the Z state. In this way, the host–photosensitizer complex converts photon energy into chemical energy in the form of out-of-equilibrium photostationary states, including ones that cannot be accessed through direct photoexcitation. (18, 19). For example, deep-sea fishes install a chlorophyll antenna next to the opsin-bound retinal (20, 21). This antenna captures red light and sensitizes the nearby retinal by means of a triplet-energy transfer (TET) mechanism (22). The subsequent photoisomerization of retinal induces a large conformational change in the surrounding opsin protein, ultimately enabling the fish to detect red light (23). Similar to retinal, azobenzene can be switched through TET (24–26). Unfortunately, this process has long (27, 28) been known to unidirectionally convert E–Z mixtures to the thermodynamically stable E isomer. This directionality originates from i) the higher tendency of Z (over E) to act as a triplet-energy acceptor (29) and ii) the preferential [by a factor of >50 (28)] relaxa- tion of the triplet excited state of azobenzene to E over Z. Therefore, whereas various photo- sensitizers can rapidly and efficiently facilitate the equilibration of the high-energy Z isomer into the stable E state, the reverse reaction— i.e., sensitized disequilibration—is far more challenging and has remained elusive. The concept of disequilibration by sensitization under confinement We hypothesized that sensitized disequilibration might be achieved by using a photosensitizer (PS) that acts on the E isomer of azobenzene with high selectivity (Fig. 1B). We have previously shown that (i) the water-soluble, palladium- containing macrocyclic host H (30) (Fig. 1C) binds two molecules of various E-azoarenes (which are planar and readily stack on top of A zobenzene and its derivatives are arguably the simplest and most widely studied photoswitchable compounds (1–3). Upon exposure to ultraviolet (UV) light, the pla- nar (4), nonpolar E isomer of azobenzene isomerizes to the metastable Z form (Fig. 1A), which is nonplanar and substantially more polar. The Z→E back-isomerization occurs spontaneously and can be accelerated with visible (blue) light. Owing to the highly re- versible nature of E⇄Z photoswitching, azo- benzenes and other azoarenes (5) have found applications in energy storage systems (6, 7), switchable catalysis (8, 9), controlled release (10, 11), and photopharmacology (12, 13), to name but a few (14, 15). However, the necessity to rely on UV light to generate the metastable Z isomer has severely limited the applicability of these compounds. Shifting E-azobenzene’s absorption band to the visible range can be achieved by decorating it with various sub- stituents (16, 17), but this approach requires additional synthetic effort and affects the com- pound’s identity. Natural systems evolved an alternative, supra- molecular strategy to extend the absorption spectral range of photoswitchable molecules 1Department of Molecular Chemistry and Materials Science, Weizmann Institute of Science, Rehovot 7610001, Israel. 2Fritz Haber Center for Molecular Dynamics Research, Institute of Chemistry, The Hebrew University of Jerusalem, Jerusalem 9190401, Israel. 3Faculty of Engineering and Natural Sciences, Tampere University, P.O. Box 541, 33101 Tampere, Finland. 4Department of Chemistry, University of Wrocław, 14 F. Joliot-Curie St., 50383 Wrocław, Poland. 5Department of Chemistry, Molecular Sciences Research Hub, White City Campus, Imperial College London, 82 Wood Lane, London W12 7SL, UK. 6Department of Chemical Research Support, Weizmann Institute of Science, Rehovot 76100, Israel. 7Institute of Science and Technology Austria, Am Campus 1, 3400 Klosterneuburg, Austria. *Corresponding author. Email: rafal.klajn@ista.ac.at (R.K.); igor.schapiro@mail.huji.ac.il (I.S.) Fig. 1. Disequilibration by sensitization under confinement (DESC). (A) The transformation of the stable E isomer of azobenzene to the metastable Z isomer traditionally relies on the use of ultraviolet (l ≈ 350 nm) light. (B) The mechanism of DESC is as follows: (i) formation of the ternary inclusion complex (E·PS)⊂H (PS, photosensitizer; H, host); (ii) absorption of a photon of visible light by the PS followed by intersystem crossing (ISC); (iii) triplet-energy transfer (TET) and the formation of triplet azobenzene, followed by its relaxation (iv) to Z-azobenzene or (iv') back to E-azobenzene (corresponding to internal conversion); and (v) disassembly of the unstable (Z·PS)⊂H inclusion complex. (C) Components of the supramolecular system used for DESC include macrocyclic host H coassembled from six Pd2+ ions and four triimidazole ligands, and a photosensitizer (e.g., BODIPY ps1). (D) Structural formulae of azoarenes 1 to 9 investigated in this study. Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E each other to form noncovalent homodimers), but only one molecule in the Z configuration (31, 32) [because of its nonplanar (33) geom- etry]; (ii) host H can also encapsulate—and thus, induce noncovalent dimerization of—guests structurally similar to E-azobenzene (i.e., planar aromatic molecules), including various dyes (34, 35); and (iii) mixing two different inclusion complexes (each binding two molecules of a given guest) induces a rapid guest exchange between the hosts, affording heterodimeric complexes, whereby the host encapsulates two different guest molecules (36). Taken together, we speculated that host H could coencapsu- late the E isomer of azobenzene and a PS (thus bringing them in close proximity) while pro- hibiting close encounters of the same PS with the Z-azobenzene (which is bound as a sole guest). We call this approach disequilibration by sensitization under confinement (DESC). The concept of DESC is illustrated in Fig. 1B. The addition of an encapsulated PS [i.e., (PS)2⊂H] to E-azobenzene induces the formation of a ternary complex (E⋅PS)⊂H (step i). Upon expo- sure to visible light, PS is promoted to a singlet excited state, which relaxes to a triplet state through intersystem crossing (ISC) (step ii). In step iii, PS transfers its triplet energy to the co- confined E-azobenzene. The resulting triplet azobenzene—which cannot be generated by di- rect photoexcitation—can either decay to the initial E isomer (step iv′) or transform into the Z state (step iv). The former case regenerates (E⋅PS)⊂H, which can be reexcited. By contrast, the latter case results in (Z⋅PS)⊂H, which is an unstable complex because Z-azobenzene is too bulky to coexist with the PS inside H. At the same time, Z as a sole guest is bound relatively weakly (fig. S103); hence, it is expelled from the host and effectively removed from the equi- librium. Thus, the azobenzene-free inclusion complex of the PS is regenerated (step v) and available for transforming additional mole- cules of E- into Z-azobenzene. To verify our hypothesis, we initially focused on the parent azobenzene E-1 and the proto- typical boron–dipyrromethene (BODIPY) dye ps1 (Fig. 1, C and D). Both E-1 and ps1 form homodimers within H’s cavity, as previously elucidated by several techniques including x-ray diffraction, nuclear magnetic resonance (NMR), and UV-visible (vis) absorption spec- troscopy (31, 32, 34). Figure 2B shows the UV- vis spectrum (dotted brown line) obtained after mixing aqueous solutions of the two homo- dimers, (E-1)2⊂H and (ps1)2⊂H, in a 1:1 molar ratio. The absorption profile in the visible range is practically identical to that of pure (ps1)2⊂H (blue dotted line), indicating a minute fraction of the (E-1⋅ps1)⊂H heterodimer (i.e., the equi- librium shown in Fig. 2A heavily favors the two homodimers). However, exposing this so- lution to low-intensity green light (wave- length at maximum intensity, lmax = 525 nm; 2.5 mW cm–2) resulted in a substantial (by ~35%) decrease of absorption in the near-UV region (Fig. 2B), indicating the E→Z isom- erization of 1. This result suggests that the small amount of (E-1⋅ps1)⊂H in equilibrium with the homodimers absorbs green light, the energy of which is eventually used to generate a large amount of the metastable Z isomer (Fig. 1B). The low illumination intensities used in our studies exclude the possibility of two-photon isomerization (37, 38), which we confirmed di- rectly through power-dependence experiments (fig. S108). To determine the scope of DESC, we extended our studies to a diverse portfolio of azobenzenes and other azoarenes, including derivatives with charged groups and electron-donating and -withdrawing substituents (Fig. 1D, 2 to 9). All of these compounds were encapsulated as homodimers within host H, which was con- firmed by NMR spectroscopy (supplementary materials). Similar to 1, most of these guests preferably existed as E2⊂H homodimers even in the presence of excess (ps1)2⊂H. However, azobispyrazole 9 (39) [and, to some extent, azopyrazole 8 (40)] showed a strong tendency to form a heterodimer with ps1, as manifested by the intense 509-nm peak in the absorption spectrum (Fig. 2C, dotted brown line). The high fraction of the heterodimer allowed us to grow single crystals and determine the struc- ture by x-ray diffraction, revealing E-9 and ps1 bound tightly inside the cavity of the host (Fig. 2D). Exposure of (E-9⋅ps1)⊂H to 525-nm light quenched its near-UV absorption, consistent with the E→Z isomerization (Fig. 2C). The putative (Z-9⋅ps1)⊂H heterodimer is unstable, forcing ps1 into homodimers, which explains why the 400- to 600-nm portion of the spec- trum at the end of the reaction is nearly iden- tical to that of pure (ps1)2⊂H (Fig. 2C). Similar to 1 and 9, compounds 2 through 8 also switched to their Z isomers when exposed to 525-nm light in the presence of (ps1)2⊂H (fig. S79). The vastly different heterodimer populations in 1 + ps1 versus 9 + ps1 mixtures do not trans- late into major differences in the reaction kinet- ics: The former comprises only ~2% heterodimer but requires only twice as much time as the latter (which has a heterodimer fraction of ~80%) to reach a photostationary state (PSS). This finding reflects the rapid guest-exchange kinetics between hosts (36), which led us to hypothesize that DESC should work efficiently also with catalytic amounts of the PS. Indeed, decreasing the amount of (ps1)2⊂H to only 0.05 equiv. with respect to (E-9)2⊂H extended the time required to reach the PSS fourfold (Fig. 2E), but did not markedly affect its composition. The finding that E-9 and ps1 form a het- erodimer in a near-quantitative yield allowed us to determine the quantum yield (QY) of DESC for this pair. Here, we note that ps1 within (E-9⋅ps1)⊂H is highly emissive, but its fluores- cence in (ps1)2⊂H is largely quenched (34). Therefore, exposing (E-9⋅ps1)⊂H to a 515-nm pulsed laser led to a gradual decrease of emis- sion (Fig. 2F, empty markers). When the ex- periment was repeated in the presence of an extra 2 and 4 equiv. of (E-9)2⊂H, however, we observed lag periods of ~8 and ~18 min, re- spectively. The stable fluorescence, despite the ongoing E→Z isomerization of 9, indicates that the concentration of (E-9⋅ps1)⊂H remains steady, which confirms the rapid exchange kinetics in our system: as soon as the isom- erized Z-9 is expelled from the host, it is re- placed by another copy of E-9 (if available). By assessing the mean number of absorbed photons required to convert the excess of E-9, we found that each successful E→Z isomeriza- tion event requires 17 photons on average (de- rivation can be found in the supplementary materials), which corresponds to a QY of ~6%, a notably high value, given the number of steps separating the excitation of ps1 from the for- mation of Z-9 (Fig. 1B). Once generated by DESC, the Z isomer can be back-isomerized to E through direct excita- tion with blue light (435 nm), and the process can be repeated for many cycles. To demon- strate the robustness of DESC, we subjected compound 9 to >100 switching cycles and did not observe any noticeable fatigue: both 9 and ps1 retained their initial absorbance values (Fig. 2G). Time-resolved spectroscopic and computational studies of DESC In solution, BODIPY dyes as simple as ps1 are poor triplet sensitizers (41, 42); therefore, the finding that ps1 acts as an efficient photo- sensitizer in DESC is unexpected. To obtain mechanistic insights into DESC, we performed transient absorption spectroscopy (TAS) and computational studies (Fig. 3). First, we studied the photoinduced dynamics of the (ps1)2⊂H homodimer using femtosecond TAS (fs-TAS). Figure 3A shows the fs absorption changes at two different wavelengths following excitation of (ps1)2⊂H with a 500-nm laser. The initial (<1 ps) ground-state bleach at 483 nm accom- panied by excited-state absorption at 412 nm is indicative of the transition from the ground state (S0) to the singlet excited state (S1) in ps1. At delay times >100 ps, the 412-nm absorp- tion increases further and the 483-nm bleach becomes more pronounced, which can be at- tributed (43) to ISC from the S1 state to the triplet excited state (T1). Using microsecond TAS (ms-TAS), we found the resulting triplet state to be remarkably stable, with a mono- exponential lifetime of 16.5 ± 0.5 ms under ambient conditions (fig. S114A). As expected from a triplet state, the lifetime is strongly de- pendent on the amount of oxygen in the sol- vent; decreasing the amount of O2 by bubbling Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Following DESC by steady-state absorption and emission spectros- copy. (A) Equilibrium between (top) homodimeric inclusion complexes of a PS and an E-azoarene—(PS)2⊂H and E2⊂H, respectively—and (bottom) the heterodimeric complex (E·PS)⊂H. Only E residing within the heterodimer, but not the homodimer, can be switched with visible light. The resulting Z is encapsulated as a sole guest (if enough host is available) and cannot be sensitized either. (B) Absorption spectrum of a 1:1 mixture of (ps1)2⊂H and (E-1)2⊂H (dotted brown line) and changes in the spectra accompanying irradiation with green light (lmax = 525 nm, denoted by green shading). The dotted blue line represents the absorption spectrum of (ps1)2⊂H. (C) Absorption spectrum of a 1:1 mixture of (ps1)2⊂H and (E-9)2⊂H [dotted brown line; predominantly (E-9·ps1)⊂H] and changes in the spectra accompanying irradiation with green light (lmax = 525 nm, denoted by green shading). The dotted blue line represents the absorption spectrum of (ps1)2⊂H. (D) The x-ray crystal structure of the heterodimeric complex (E-9·ps1)⊂H (from the left: front view, side view, and top view; light gray, host H; dark gray, E-9; green, ps1; water molecules, counterions, and host protons omitted for clarity). (E) Graph following DESC of 9 in the presence of different equiv. of ps1 (the data were normalized to the 0 to 1 range, except the experiment with no PS; raw spectra can be found in fig. S76). Abs., absorbance. (F) Evolution of the emission intensity of ps1 under 515-nm light (used both to induce DESC and excite fluorescence) as a function of the amount of 9. Em. int., emission intensity. (G) More than 100 cycles of reversible photoisomerization of 9 induced solely by visible light (E→Z, DESC with 525-nm light for 2 min; Z→E, direct photoexcitation using 435-nm light for 30 s). The amount of the E isomer is proportional to absorbance at 353 nm; the absorption at 480 nm originates from the (ps1)2⊂H homodimer. norm., normalized. N2 for 4 min and 10 min extended the lifetime of the T1 state of ps1 to 160 ± 4 ms and 10.1 ± 0.5 ms, respectively (fig. S114B). When the fs-TAS experiment was repeated for a 1:2 mixture of (ps1)2⊂H and (E-9)2⊂H (i.e., a pair with a high tendency to form a hetero- dimer), the bleach at 483 nm was substantially less pronounced (Fig. 3B, inset), indicating a TET to E-9 (Fig. 1B, step iii). Notably, the TET and the subsequent formation of Z-9 occur in the nanosecond time regime—i.e., much faster than the lifetime of the ps1 triplet state—which explains why DESC does not require exclusion of oxygen. In fact, we found the process to be equally efficient in strictly deoxygenated ver- sus thoroughly oxygenated water (fig. S91). We also studied the E-9–ps1 pair under ambient conditions by ms-TAS (Fig. 3C) and found the intensity of transient absorption at 430 nm within 0.1 ms after excitation (DAbs*430) to be inversely proportional to the amount of E-9. This finding was consistent with the quench- ing of ps1’s triplet state by its E-9 co-guest by means of TET. The resulting triplet-9 can either relax to the initial E-9 isomer or switch to Z-9, which absorbs at 430 nm, hence the increasing steady-state absorption (DAbs∞ 430). This intimate relationship between the degree of ps1 triplet state quenching and the extent of E→Z isom- erization identified by ms-TAS further confirms that DESC proceeds by means of TET between the ps1 donor and E-9 acceptor. To gain further insights into DESC, we studied various azoarene–PS combinations as nonco- valent heterodimers by using quantum chem- ical simulations. We consistently found that the lowest-energy triplet state within these heterodimers was localized on the PS (there- fore, we refer to it as TPS) and the second-lowest triplet state was localized on the azoarene (i.e., Tazo), which indicates that the TPS→Tazo tran- sition is an endothermic process (25, 44) (e.g., Fig. 3D shows the energy diagram for 1⋅ps1. These results led us to hypothesize that the experimentally observed TET might be facili- tated by thermal fluctuations of molecules, as suggested previously for other triplet donor– acceptor pairs (25, 26, 45, 46). Therefore, we Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Time-resolved spectroscopic and computational studies of DESC. (A) Normalized decays of fs transient absorption of ps1 at 412 nm and 483 nm within the (ps1)2⊂H homodimer (excitation wavelength, lexc = 500 nm). Points, raw data; lines, four-exponential fits (details can be found in the supplementary materials). norm., normalized. (B) Normalized transient-absorption decays of ps1 at 483 nm (lexc = 500 nm) in (ps1)2⊂H versus the (E-9·ps1)⊂H heterodimer (linear scale in the 0 to 1 ps time range; logarithmic scale beyond 1 ps). (Inset) The same data plotted on the linear scale. (C) Decays of ms transient absorption at 430 nm (lexc = 510 nm) in (ps1)2⊂H in the presence of increasing amounts of (E-9)2⊂H. The thin and thick lines correspond to experimental data and biexponential fits, respectively. (Inset) The inverse correlation between DAbs*430 (absorbance at 430 nm immediately after photoexcitation) and DAbs∞ (steady-state absorbance after photoswitching). All the TAS results presented here were collected under ambient (nondeoxygenated) conditions. (D) The calculated energies of the E-1·ps1 heterodimer's lowest excited states: the bright singlet state (Sps1) and the two lowest triplet states, localized on ps1 and 430 E-1 (Tps1 and TE-1, respectively) (details can be found in fig. S116). The arrows indicate the sequence of events (“hn” indicates a photoinduced transition; the gray arrow indicates an endothermic process). (E) Ground-state relaxed scan along the C–N=N–C dihedral angle F in 1 within the 1·ps1 heterodimer. The gray line denotes the ground state (S0), the red line denotes the Sps1 state, and the blue lines denote the TE-1 and Tps1 states. The circular and triangular markers correspond to the localization of the excited state on the donor (ps1) and acceptor (E-1), respectively. (F) Ground-state relaxed scan of F in 1 within the (1·ps1)⊂H heterodimer (orange trace). The blue trace shows the energies that correspond to the same configurations of 1 and ps1 after removing the host and its interactions. (i), (ii), and (iii) correspond to DF values 0°, +165°, and –170°, respectively. (G) Optimized geometries of (1·ps1)⊂H for the three DF values indicated in (F) (left, side views; right, top views). The distances between the indicated equatorial Pd nodes describe the degree of host deformation; the larger the difference between the two Pd–Pd distances, the greater the transition of H from a tube-like conformation into a bowl-like conformation. studied the dependence of the 1⋅ps1 excited- state energies on the C–N=N–C dihedral angle (F) in azobenzene 1 (which is substantially more flexible than ps1). Figure 3E shows a relaxed scan for the 1⋅ps1 heterodimer, demon- strating that an 18° twist in F is sufficient to invert the energetic order of Tps1 and TE-1, making TET energetically favorable. We sep- arately studied the dynamics of the (1⋅ps1)⊂H heterodimer by multiscale molecular dynamics simulations (fig. S121 and movie S1). These simulations revealed that thermal fluctuations readily allow 1 to adopt conformations with DF ≥ 18° at room temperature. We performed additional multiscale simu- lations to better understand the relative insta- bility of the (Z⋅PS)⊂H heterodimers versus (E⋅PS)⊂H (Fig. 1B), which lies at the heart of efficient DESC. The starting point of the sim- ulations was (E-1⋅ps1)⊂H with a perfectly planar geometry of E-1. We performed a relaxed scan by changing F in steps of 5° in both senses of rotation; the resulting energies are plotted in Fig. 3F in orange. The blue curves in Fig. 3F cor- respond to the same geometries while neglecting the host and its interactions; therefore, the energetic difference between the two curves (highlighted with gray shading) quantifies the instability of the inclusion complex. We found that rotating F in one direction affords a highly Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E unstable supramolecular architecture (Fig. 3G, ii) that is ~0.35 eV (~8 kcal/mol) higher in energy than free Z-1⋅ps1 (Fig. 3F). Notably, rotating F in the opposite direction gave rise to a geometry where the host did not markedly increase the energy (Fig. 3, F and G, iii). However, in this struc- ture, host H assumes a bowl-like conformation, and Z-1 extrudes from the cavity, facilitating its expulsion to the solution (movies S5 and S6). The high conformational flexibility of H (47) is an important requirement for DESC; indeed, the cavities of rigid coordination cages (48) and other confined environments (49, 50) were shown to render azobenzene nonphotoswitchable under all wavelengths of light. Tuning the excitation wavelength of DESC Encouraged by the unexpected sensitization potency of ps1 under confinement, we consid- ered DESC with other, more red-shifted dyes, including ones not previously known to act as triplet sensitizers. To this end, we first focused on the fluorinated BODIPY ps2 (Fig. 4A), with an absorption peak centered at 553 nm (51) (compared with 499 nm for ps1). We found that ps2 exhibited a higher affinity than ps1 to form heterodimers with various azoarenes and hypothesized that the increased PS–azoarene interactions should further promote DESC. Indeed, Fig. 4B shows that ps2 induces a near- quantitative E→Z conversion of an equimo- lar amount of azobenzene 4 within only 90 s of low-intensity (2.5 mW cm–2) yellow-light (561 nm) irradiation. The more efficient DESC allowed us to decrease the PS loading further: At only 0.01 equiv. of (ps2)2⊂H with respect to (E-4)2⊂H, the PSS was reached within ~20 min (Fig. 4C). In general, ps2 is a more efficient DESC agent than ps1. However, we found one exception: Fig. 4. Extending the concept of DESC to red-shifted photosensitizers. (A) Structural formulae of fluorinated BODIPY ps2, resorufin ps3, and resazurin ps4. (B) Changes in the absorption spectra of encapsulated E-4 in the presence of an equimolar amount of encapsulated sensitizer ps2 under yellow light (lmax = 561 nm; 2.5 mW cm–2). (C) DESC (here, for E-4) in the presence of substoichiometric amounts of ps2. (D) Changes in the absorption spectra of encapsulated E-1 in the presence of an equimolar quantity of encapsulated sensitizer ps3 under orange light (lmax = 599 nm; 0.8 mW cm–2). (E) DESC (here, for E-1) in the presence of substoichiometric amounts of ps3. (F) Changes in the absorption spectra of encapsulated E-2 in the presence of an equimolar quantity of encapsulated sensitizer ps4 under red light (lmax = 635 nm; 3.4 mW cm–2). (G) DESC (here, for E-2) in the presence of substoichiometric amounts of ps4. The data in (C), (E), and (G) were normalized to the 0 to 1 range, except for the experiments with no PS; raw data can be seen in figs. S85G, S93F, and S98G, respectively. norm., normalized. ps2 proved unable to induce the switching of azobispyrazole E-9. To understand this re- sult, we resorted to quantum chemical simu- lations and found the TPS–Tazo energy gap for the E-9–ps2 heterodimer to be exceptionally high (1.03 eV; compared with 0.26 eV for E-1–ps1 in Fig. 3D). Relaxed scans analogous to those in Fig. 3E showed that F in 9 must twist by 38°—a prohibitively large distortion—for the energies of these two triplet states to equalize (fig. S122D). These computational results not only rationalize the experimental findings but also provide further (although indirect) sup- port for the involvement of the TET mechanism in our system. We also worked with resorufin ps3 and res- azurin ps4 (Fig. 4A), both of which were pre- viously reported to form inclusion complexes of the (PS)2⊂H type (35). These two dyes are red-shifted even further than is ps2; for ex- ample, the absorption maxima of their respec- tive heterodimers with E-1 appear at 587 and 616 nm, with absorption extending into the red spectral range. To our satisfaction, exciting the absorption bands on these heterodimers with orange and red light, respectively, resulted in a highly efficient E→Z isomerization of nearly all azoarene–PS combinations (Fig. 4 and figs. S95 and S100). The performance of DESC is showcased in Fig. 5A, which lists the PSS compositions (blue font) for all the nine model azoarenes shown in Fig. 1D (encapsulated within H in water with 0.05 equiv. of the selected sensitizer: ps2 for 1 to 7 and ps1 for 8 and 9). The reactions were performed on the NMR scale (i.e., mil- ligram quantities of 1 to 9) and can readily be scaled up to obtain the Z isomers on the pre- parative scale (tens of milligrams). As control experiments (red font), we irradiated 1 to 9 under the same conditions and in the pres- ence of the PS, but without host H (hence, in an organic solvent). In the absence of H, the E isomers could not be co-confined with the PS, which resulted in negligible amounts of Z- isomer formation by direct photoexcitation [only azobenzene 6, known for its visible- light-responsiveness (17), afforded a small (14%) amount of Z]. The positively charged azobenzene (52) 5 [often recognized as a proto- typical photopharmacophore (13, 53)] showed a particularly impressive contrast in behavior between the presence and absence of the host, giving rise to 98% of Z. Notably, such a Z-rich PSS cannot be achieved by direct photoisom- erization [of neither (E-5)2⊂H nor free 5] with any wavelength of light (the same is true for compounds 1, 3, and 6) because of the partial overlap of the absorption bands of the two isomers (54). By contrast, the PSS composition in DESC is dictated by the tendency of the two isomers to form the ternary (azo⋅PS)⊂H com- plex, and this tendency is overwhelmingly higher for the E isomer. Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. Performance and selectivity of DESC. (A) Numbers in blue represent the composition of the PSS of azoarenes 1 to 9 subjected to DESC on the millimolar scale in the presence of 0.05 equiv. of the PS (ps2 under 561-nm light for 1 to 7; ps1 under 525-nm light for 8 and 9). Numbers in red represent the PSS compositions of the same azoarene–PS mixtures under identical illumination conditions, but in the absence of the host (CDCl3 was used as the solvent for all azoarenes except 3 and 5, for which CD3OD was used). Illumination times were 15, 12, 35, 12, 40, 24, 45, 4, and 9 min for 1 to 9, respectively. n.d., not detected (i.e., the amount of Z was below the NMR detection limit). (B) Evolution of absorption spectra during red-light irradiation of E-3 in water in the presence of 0.005 equiv. of (ps4)2⊂H. The resulting PSS contains 88% Z-3. (C) Selective switching of the negatively charged E-3 with 561-nm light in the presence of the positively charged E-5 on the micromolar scale. (D) UV-vis absorption spectra for the experiment shown in (C). Dotted line, PSS under 561-nm light; dashed line, PSS after the subsequent exposure to 365-nm light (where both azobenzenes isomerize to a similar extent). (E) Selective switching of E-4 with 561-nm light and (ps2)2⊂H in the presence of a UV-dimerizable anthracene. (F) UV-vis absorption spectra for the experiment shown in (E). Dotted line, PSS under 561-nm light; dashed line, after the subsequent exposure to 365-nm light. Having demonstrated that the positively charged E-5 can be successfully transformed into Z-5 despite its low affinity to the like- charged H, we speculated that other water- soluble azobenzenes may also be efficiently disequilibrated with a substoichiometric amount of not only the PS but also the host. Figure 5B shows the result of an experiment in which an aqueous solution of the negatively charged 3 was exposed to red light in the presence of 0.005 equiv. of (ps4)2⊂H. The absorption spec- trum of this solution is dominated by the intense absorption peak of E-3 in the near-UV region; the small amount of the sensitizer appears as a weak band at ~600 nm (Fig. 5B). Remarkably, exciting this band with low-intensity 635-nm light resulted in a near-complete disappearance of the much more prominent and distant peak originating from another species (E-3). We found that the PSS contained 88% of Z-3 (versus ~0% in the absence of either H or ps4), which indi- cated that each molecule of H hosted more than 180 E→Z isomerization events on average. Photoswitching selectivity enabled by DESC To further demonstrate the potential of DESC, we explored the charge (+12) and cavity size of host H to discriminate between photoreac- tive compounds with overlapping absorption bands, which otherwise cannot be converted selectively. To this end, we mixed E-3 and E-5 in a 1:3 ratio and added (ps2)2⊂H (0.5 equiv. with respect to 3) (Fig. 5C). At low (micromo- lar) concentrations, only the negatively charged 3 exhibits affinity to H; 5 is not encapsulated owing to the Coulombic repulsion. Indeed, yellow-light (561 nm) illumination of this mix- ture led to a highly selective switching of E-3 (despite the threefold excess of E-5) (Fig. 5D and fig. S104). By contrast, exposure to UV light induced nonselective isomerization of both azobenzenes by direct excitation. In the sec- ond example, we worked with a mixture of E-4 and 9-bromoanthracene, both in the form of homodimers encapsulated within H. Upon exposure to UV light, the encapsulated an- thracene rapidly dimerizes to afford the cor- responding dianthracene (36) under the same irradiation conditions that triggered the direct E→Z photoisomerization of 4 (Fig. 5E, dashed arrow). However, exposing the same mixture Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E to yellow light in the presence of (ps2)2⊂H induced highly selective photoisomerization of azobenzene, leaving the anthracene intact (Fig. 5F, dotted line). Discussion From a thermodynamic perspective, our sys- tem acts as a light-driven supramolecular ma- chine that converts light into chemical energy in the form of out-of-equilibrium photosta- tionary states. DESC relies on the selective coencapsulation of the stable E isomer of an azobenzene with a dye that acts as an antenna, absorbing visible light energy that is ultimately used to generate the metastable Z isomer. The absorption of light promotes the dye from the ground state to the singlet excited state. Con- finement inside of the host increases the dye’s ability to undergo intersystem crossing, pop- ulating the dye’s triplet state and turning it into a potential triplet sensitizer. Quantum chemical simulations reveal that although the triplet state of azobenzene is higher than that of the dye, a small dihedral angle twist in azobenzene lowers its triplet energy while in- creasing the triplet energy of the coencapsu- lated dye to the extent that the two energy levels converge. Therefore, the dye-to-azobenzene triplet-energy transfer can become favorable owing to azobenzene dynamics (25, 44). Once in the triplet state, the azobenzene can either dissipate energy or switch to the Z isomer. Z-azobenzene is nonplanar and can no longer be co-confined with the photosensitizer; thus, it is expelled from the host and cannot be re- sensitized. In this way, DESC shifts the equi- librium toward the metastable Z state without the need to populate azobenzene’s singlet ex- cited state, which is relatively high in energy and requires the absorption of UV light. Although we focused on a particular host and one class of photoswitchable molecules (H and azoarenes, respectively), our results allow us to establish general design principles for other DESC systems: (i) The host should have an affinity for a photoswitch and a photosen- sitizer, and its cavity should be large enough to simultaneously encapsulate the photosensitizer and the thermodynamically stable isomer of the photoswitch; (ii) the host’s affinity for the metastable form of the photoswitch must be substantially lower (because of its different shape and/or polarity) so as to render coencap- sulation with the photosensitizer unfavorable; and (iii) an open cavity and/or conformational flexibility should promote rapid guest-binding and -release kinetics by the host for fast eatalytic turnover and to ensure that once generated, the metastable form of the photoswitch is expelled from the cavity before it is resensitized. In principle, DESC is applicable to other classes of photoswitchable compounds, although larger or differently shaped hosts may be required. Self-assembly through metal–ligand coordination offers an attractive approach to generating a wide range of hosts from simple components in a modular fashion. As demonstrated in this work, DESC is a ro- bust process that works with catalytic amounts of sensitizers, under ambient conditions (no oxygen exclusion necessary), and for a wide range of azobenzenes and heterocyclic azoarenes. We envision that DESC will become a powerful tool to control chemical reactivity through a com- bination of light irradiation and confinement. RE FERENCES AND NOTES 1. G. S. Hartley, Nature 140, 281 (1937). 2. H. M. D. Bandara, S. C. Burdette, Chem. Soc. Rev. 41, 1809–1825 (2012). 3. F. A. Jerca, V. V. Jerca, R. Hoogenboom, Nat. Rev. Chem. 6, 51–69 (2022). 4. T. Tsuji, H. Takashima, H. Takeuchi, T. Egawa, S. Konaka, J. Phys. Chem. A 105, 9347–9353 (2001). 5. S. Crespi, N. A. Simeth, B. König, Nat. Rev. Chem. 3, 133–146 (2019). 6. Z. Wang et al., Joule 5, 3116–3136 (2021). 7. M. Le, G. G. D. Han, Acc. Mater. Res. 3, 634–643 (2022). 8. R. S. Stoll, S. Hecht, Angew. Chem. Int. Ed. 49, 5054–5075 (2010). 9. Y. Wei, S. Han, J. Kim, S. Soh, B. A. Grzybowski, J. Am. Chem. Soc. 132, 11018–11020 (2010). 10. J. Lu, E. Choi, F. Tamanoi, J. I. Zink, Small 4, 421–426 (2008). 11. D. Tarn et al., Nanoscale 6, 3335–3343 (2014). 12. W. A. Velema, W. Szymanski, B. L. Feringa, J. Am. Chem. Soc. 136, 2178–2191 (2014). 13. K. Hüll, J. Morstein, D. Trauner, Chem. Rev. 118, 10710–10747 (2018). 14. J. M. Mativetsky et al., J. Am. Chem. Soc. 130, 9192–9193 (2008). 15. R. Klajn, P. J. Wesson, K. J. M. Bishop, B. A. Grzybowski, Angew. Chem. Int. Ed. 48, 7035–7039 (2009). 16. A. A. Beharry, O. Sadovski, G. A. Woolley, J. Am. Chem. Soc. 133, 19684–19687 (2011). 17. D. Bléger, J. Schwarz, A. M. Brouwer, S. Hecht, J. Am. Chem. Soc. 134, 20597–20600 (2012). 18. S. P. Balashov et al., Science 309, 2061–2064 (2005). 19. A. Chazan et al., Nature 615, 535–540 (2023). 20. J. K. Bowmaker, Vision Res. 48, 2022–2041 (2008). 21. T. Isayama et al., Nature 443, 649–649 (2006). 22. A. Valentini, M. Nucci, L. M. Frutos, M. Marazzi, ChemPhotoChem 3, 925–932 (2019). 23. R. H. Douglas et al., Nature 393, 423–424 (1998). 24. J. Moreno, M. Gerecke, L. Grubert, S. A. Kovalenko, S. Hecht, Angew. Chem. Int. Ed. 55, 1544–1547 (2016). 25. J. Isokuortti et al., Chem. Sci. 12, 7504–7509 (2021). 26. P. Bharmoria et al., Chem. Sci. 13, 11904–11911 (2022). 27. L. B. Jones, G. S. Hammond, J. Am. Chem. Soc. 87, 4219–4220 (1965). 28. E. Fischer, J. Am. Chem. Soc. 90, 796–797 (1968). 29. F. Strieth-Kalthoff, M. J. James, M. Teders, L. Pitzer, F. Glorius, Chem. Soc. Rev. 47, 7190–7202 (2018). 30. D. Samanta, S. Mukherjee, Y. P. Patil, P. S. Mukherjee, Chem. Eur. J. 18, 12322–12329 (2012). 31. D. Samanta et al., Proc. Natl. Acad. Sci. U.S.A. 115, 9379–9384 (2018). 32. A. I. Hanopolskyi et al., Beilstein J. Org. Chem. 15, 2398–2407 34. J. Gemen, J. Ahrens, L. J. W. Shimon, R. Klajn, J. Am. Chem. Soc. 142, 17721–17729 (2020). 35. O. Yanshyna, M. J. Białek, O. V. Chashchikhin, R. Klajn, Commun. Chem. 5, 44 (2022). 36. J. Gemen et al., Chem 8, 2362–2379 (2022). 37. H. Ishitobi, Z. Sekkat, S. Kawata, J. Chem. Phys. 125, 164718 (2006). 38. C.-K. Lim et al., Nanoscale 8, 4194–4202 (2016). 39. A. Gonzalez et al., J. Am. Chem. Soc. 144, 19430–19436 (2022). 40. C. E. Weston, R. D. Richardson, P. R. Haycock, A. J. P. White, M. J. Fuchter, J. Am. Chem. Soc. 136, 11878–11881 (2014). 41. M. A. Filatov, Org. Biomol. Chem. 18, 10–27 (2019). 42. I. Kähärä et al., Photochem. Photobiol. Sci. 21, 1677–1687 (2022). 43. R. P. Sabatini et al., J. Phys. Chem. Lett. 2, 223–227 (2011). 44. K. Sandros et al., Acta Chem. Scand. 16, 958–968 (1962). 45. J. Eng, T. J. Penfold, Chem. Rec. 20, 831–856 (2020). 46. S. A. Mavrommati, S. S. Skourtis, J. Phys. Chem. Lett. 13, 9679–9687 (2022). 47. L. Pesce, C. Perego, A. B. Grommet, R. Klajn, G. M. Pavan, J. Am. Chem. Soc. 142, 9792–9802 (2020). 48. T. Kusukawa, M. Fujita, J. Am. Chem. Soc. 121, 1397–1398 (1999). 49. A. B. Grommet, L. M. Lee, R. Klajn, Acc. Chem. Res. 53, 2600–2610 (2020). 50. R. Wang et al., J. Electroanal. Chem. 438, 213–219 (1997). 51. S. Choi, J. Bouffard, Y. Kim, Chem. Sci. 5, 751–755 (2014). 52. J. Bieth, S. M. Vratsanos, N. Wassermann, B. F. Erlanger, Proc. Natl. Acad. Sci. U.S.A. 64, 1103–1106 (1969). 53. M. M. Lerch, M. J. Hansen, G. M. van Dam, W. Szymanski, B. L. Feringa, Angew. Chem. Int. Ed. 55, 10978–10999 (2016). 54. E. Fischer, J. Phys. Chem. 71, 3704–3706 (1967). 55. J. Gemen et al., Disequilibrating azobenzenes by visible-light sensitization under confinement, Dryad (2023); https://dx.doi. org/10.5061/dryad.4qrfj6qg1. AC KNOWLED GME NTS Funding: We acknowledge funding from the European Union’s Horizon 2020 Research and Innovation Program [European Research Council grants 820008 (Ra.K.) and 101045223 (A.P.) and Marie Skłodowska-Curie grants 812868 (J.G.) and 101022777 (T.-P.R.)], the Academy of Finland [Center of Excellence Programme LIBER grant 346107 (A.P.), Flagship Programme PREIN grant 320165 (A.P.), and Postdoctoral Researcher grant 340103 (T.-P.R.)], Zuckerman STEM Leadership Program Fellowship (J.R.C.), President’s PhD Scholarship (M.O.), and the EPSRC [Established Career Fellowship grant EP/R00188X/1 (M.J.F.)]. Author contributions: Conceptualization: A.P., I.S., and Ra.K.; Methodology: J.G., J.R.C., T.-P.R., N.D., D.O., A.P., I.S., and Ra.K.; Investigation: J.G., T.-P.R., N.D., M.J.B., M.W., M.F., M.K., M.O., V.A.B., Ri.K., and Y.D.-P.; Funding acquisition: M.J.F., A.P., I.S., and Ra.K.; Project administration: Ra.K.; Supervision: D.O., M.J.F., A.P., I.S., and Ra.K.; Writing: J.G., J.R.C., T.-P.R., N.D., A.P., I.S., and Ra.K. Competing interests: The authors declare no competing interests. Data and materials availability: Crystallographic data for inclusion complexes (E-9)2⊂H, (E-9⋅ps1)⊂H, and (TX)2⊂H have been deposited in the Cambridge Crystallographic Data Centre (CCDC) with accession codes 2240153, 2227254, and 2227255, respectively. All other data are available in the main text or the supplementary materials. Raw data underlying the plots have been deposited at Dryad (55). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adh9059 Materials and Methods Supplementary Text Figs. S1 to S122 Tables S1 to S5 References (56–93) Movies S1 to S6 (2019). 33. A. Mostad et al., Acta Chem. Scand. 25, 3561–3568 (1971). Submitted 21 March 2023; accepted 22 August 2023 10.1126/science.adh9059 Gemen et al., Science 381, 1357–1363 (2023) 22 September 2023 7 of 7
10.1126_science.adf9725
RES EARCH NEUROSCIENCE Wide-field fluorescence lifetime imaging of neuron spiking and subthreshold activity in vivo Adam J. Bowman1*, Cheng Huang2,† Mark J. Schnitzer2,3,4, Mark A. Kasevich1* The development of voltage-sensitive fluorescent probes suggests fluorescence lifetime as a promising readout for electrical activity in biological systems. Existing approaches fail to achieve the speed and sensitivity required for voltage imaging in neuroscience applications. We demonstrated that wide-field electro-optic fluorescence lifetime imaging microscopy (EO-FLIM) allows lifetime imaging at kilohertz frame-acquisition rates, spatially resolving action potential propagation and subthreshold neural activity in live adult Drosophila. Lifetime resolutions of <5 picoseconds at 1 kilohertz were achieved for single-cell voltage recordings. Lifetime readout is limited by photon shot noise, and the method provides strong rejection of motion artifacts and technical noise sources. Recordings revealed local transmembrane depolarizations, two types of spikes with distinct fluorescence lifetimes, and phase locking of spikes to an external mechanical stimulus. R ecording the electrical activity of neu- rons at high spatial and temporal reso- lution is central to understanding brain function. Fluorescent probes of cal- cium activity enable optical studies of large neuron populations in vivo (1, 2). How- ever, the response time of calcium indicators is much slower than the underlying electri- cal signals. Fluorescent voltage sensors are a complementary approach, providing direct readout of neuron membrane potential with the capability to resolve action potentials. Although voltage probes have rapidly devel- oped with a variety of genetically encoded (3–7) and chemical dyes (8) in use, there re- main considerable challenges to their appli- cation in vivo because of low sensitivity, rapid photobleaching, and motion artifacts. To achieve high-speed recording and sufficient signal-to-noise ratios (SNRs), most voltage probes use fluorescence intensity to read out an underlying sensing mechanism on the basis of absorption, Förster resonance energy trans- fer (FRET), or quenching. We applied an alternative strategy for de- tecting fast probe dynamics that is based on lifetime imaging (9). Voltage sensors that use FRET and quenching mechanisms modulate the probe’s nonradiative decay rate, intrinsi- cally connecting fluorescence intensity with nanosecond excited-state lifetime. Lifetime is a promising readout for voltage imaging, es- pecially because of its capability to provide an absolute indication of membrane potential (10). Recent results have validated fluorescence 1Physics Department, Stanford University, Stanford, CA 94305, USA. 2James H. Clark Center, Stanford University, Stanford, CA 94305, USA. 3CNC Program, Stanford University, Stanford, CA 94305, USA. 4Howard Hughes Medical Institute, Stanford University, Stanford, CA 94305, USA. *Corresponding author. Email: abowman2@stanford.edu (A.J.B.); kasevich@stanford.edu (M.A.K.) †Present address: Department of Neuroscience, Washington Univer- sity School of Medicine, St. Louis, MO 63110, USA. lifetime for static measurements of membrane potential in vitro (11). However, existing life- time detectors—for example, single-photon counters or cameras with modulated pixels (12)—fall short of the requirements for detect- ing fast dynamics and imaging neurons in vivo, either because of their limited photon through- put or because of prohibitively high noise. EO-FLIM technique We demonstrated an approach that uses electro- optic fluorescence lifetime imaging micros- copy (EO-FLIM), an all-optical technique for lifetime imaging that is based on nanosecond gating of a wide-field image (13, 14). EO-FLIM allows efficient photon collection and is com- patible with detection on high-speed, low- noise scientific cameras. With this method, we achieved lifetime imaging of action potentials in vivo. EO-FLIM enables significant suppression of intensity artifacts, allowing robust imaging in the presence of tissue motion and fluctua- tions in illumination intensity. Such artifacts are ubiquitous in recordings of neural activity from awake, behaving animals (15, 16). This has two consequences: (i) It enables faithful recording of subthreshold voltage waveforms, and (ii) it improves the SNR by suppressing high-frequency intensity noise. These follow from the fact that EO-FLIM estimates lifetime from the ratio of a pair of simultaneously recorded intensity channels derived from a common optical source. In conventional ap- proaches, corrections for intensity noise use ratios of measurements that are nonsimulta- neous and easily corrupted by high-frequency noise or slower motion artifacts. Optical sen- sors of neuron activity are typically reported by DF =F , referencing fast intensity changes (DF) to a nonsimultaneous, average fluores- cence baseline (F ). Fluorescence lifetime can read out an intensity-optimized sensor with improved temporal noise performance and long-term stability without sacrificing acqui- sition speed. We implemented our approach through in- corporation of a Pockels cell into the fluorescence detection path of a standard epifluorescence microscope (Fig. 1A, fig. S1, and materials and methods). The Pockels cell design was optimized for resonant drive and wide-field imaging, in- corporating thermal control and transverse crystal geometry to cancel off-axis birefrin- gence. A high-voltage modulation was applied to the Pockels cell with a resonant transformer (figs. S2 and S3), resulting in a fast polariza- tion rotation that was synchronous with the excitation pulses from a 100-ps laser source. Fluorescence from the sample was first po- larized, then polarization was modulated with the Pockels cell and finally split with a polariz- ing beamsplitter into two wide-field images on a scientific complementary metal-oxide semi- conductor (sCMOS) camera corresponding to gated (G) and ungated (U) intensity. These two images encoded nanosecond time infor- mation in their intensity ratio. Because gating was performed with a beamsplitter, it was pos- sible to capture the entire fluorescence decay with photon efficiency limited by optical coat- ings. In this study, we modulated one input polarization and discarded half of the avail- able signal on a first polarizer. This can be avoided in the future through addition of a second beam path (13). The fluorescence de- cay at each pixel was convolved with the tem- poral gating function of the Pockels cell and then sampled at a single modulation phase relative to the excitation laser (Fig. 1B). Each image thus provided a lifetime estimate for every pixel in parallel at the frame rate of the scientific camera. When imaging genetically targeted neurons in vivo, this technique al- lowed for 1-kHz–frame rate recordings with a lifetime sensitivity of 2.53 ± 0.48 ps with 0.7 × 107 to 1.4 × 107 detected photons per frame (Fig. 1C), indicating a substantial im- provement in throughput over previous FLIM recordings of cellular dynamics. EO-FLIM approaches fundamental sensitivity limits for estimating lifetimes between 1 and 4 ns (fig. S2). Results We used EO-FLIM to image a genetically en- coded voltage indicator (GEVI) in Drosophila melanogaster. We expressed pAce positive polarity GEVI in a subtype of fly mushroom body output neuron (MBON), MBON-g1pedc>ab. pAce works through FRET and is a fusion of the bright fluorescent protein mNeonGreen with a voltage-sensitive opsin from Acetabularia (3). We surgically prepared Drosophila before imaging to provide optical access to the brain (17, 18). Action potentials and subthreshold dynamics were readily resolved with action potentials corresponding to a 20- to 50-ps Bowman et al., Science 380, 1270–1275 (2023) 23 June 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Lifetime imaging of action potentials. (A) Schematic of EO-FLIM microscope. Wide-field fluorescence images were modulated with a Pockels cell (PC) placed between crossed polarizers (P and PBS) and driven at 20 MHz by a high-voltage resonant transformer. Two spatially offset output images were simultaneously captured after a second polarizing beamsplitter (PBS) on a sCMOS camera, corresponding to gated (G) and ungated (U) intensities. (B) Instrument response function (IRF) and fluorescence traces for the U channel were measured by varying the Pockels-cell drive phase relative to the excitation laser. The pAce GEVI was fit to 2.2-ns lifetime. For kilohertz imaging, a single optimal phase point was captured (vertical line at 0-ns delay), and the G/U image intensity ratio was converted to a lifetime estimate (fig. S1). (C) Histogram of measurements (highpass filtered) obtained at 1 kHz for a single neuron in vivo demonstrate a lifetime sensitivity of 3 ps (full trace in Fig. 2A). (D) Wide-field image of a neuron with structures indicated. Scalebar, 25 mm. (E) Whole-cell lifetime trace resolves action potentials and sub- threshold transitions. (F and G) Average spike shape is plotted in intensity and lifetime from color-coded regions. (H) Frames from an interpolated lifetime movie demonstrate spike propagation, averaging the signal from ~300 individual spikes. The point of initiation is indicated by the arrow, and bidirectional propagation was observed both along the axon and backward toward soma and dendrites (movies S1 to S3). Spike propagation was also imaged directly without averaging (movies S4 and S5). (I) Applying a 10-frame moving average allowed subthreshold signals to be localized to neuron structures (movies S6 and S7). Example frames demonstrate localization in the dendrite for both positive and negative subthreshold signals. shift in the fluorescent lifetime (Fig. 1, D and E). Average spike readouts in lifetime and in- tensity for different neuron subregions were in strong agreement both in their relative tim- ing and amplitudes (Fig. 1, F and G). The donor fluorescence lifetime of a FRET GEVI depends on radiative decay rate, kf, and the voltage-sensitive nonradiative decay rate, knr(V), associated with FRET as t ¼ 1 kf þ knrðV Þ (10, 19). In pAce, the Ace opsin acts as acceptor and provides voltage sensitivity. The donor’s fluorescence intensity is directly quenched by FRET, giving a signal, DF º sDknr, where s is the donor excitation cross section. For an ð ideal FRET process, one expects to find Dt=t ¼ DF =F. pAce gave a linear but attenuated life- ÞDF =F (fig. S4). time response of 0:70 T 0:07 This may indicate components of the GEVI response in intensity—for example, modulation of cross section s—that did not affect lifetime readout. Wide-field lifetime imaging correlates neu- ron activity with spatial structure. The point of action potential initiation in the axon is resolved along with bidirectional propagation along the axon and backward toward the den- drite and soma (20). Action potentials were attenuated and broadened in the dendrites and soma (Fig. 1, F and G, and figs. S5 and S6). Spike propagation is shown in movies S1 to S3 with still frames from movie S1 displayed in Fig. 1H, generated with spike-triggered av- eraging over ~300 spikes and interpolating between frames. We also observed individ- ual spikes and spike propagation in real time without temporal averaging (movies S4 and S5). Comparison of recordings from multiple neuronal subregions revealed local depolari- zations in the axons, which fail to initiate action potentials across the entire cell. These depolarizations were not resolved in inten- sity readout but are clearly visible in lifetime Bowman et al., Science 380, 1270–1275 (2023) 23 June 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Lifetime suppresses intensity noise and improves fidelity of subthreshold recording. Six example MBON neurons are shown comparing lifetime (blue) with DF=F intensity recordings (black). Recordings were obtained by averaging over high-resolution images shown at far left. Scalebar, 25 mm. Shaded boxes highlight some notable regions of the traces for improved lifetime readout. For each example, the distributions of spike SNR are compared for intensity and lifetime, with calculated spike detection fidelity d′ indicated. (A and B) Two examples of flies without motion demonstrate improvement of technical noise floor at high frequencies by up to 7 dB. The noise power spectra for the traces are compared at far right, with dotted lines indicating the photon shot-noise limits. (C to E) Three examples of flies having low-frequency noise associated with motion artifacts. Lifetime improves noise power spectrum across temporal frequencies, rejecting intensity noise by up to 9 dB at low frequencies (further analysis is provided in Fig. 3). (F to J) Lifetime provided an improved readout of two spike amplitudes in response to mechanical stimulus at 60 Hz. Large-amplitude (L) spikes showed an enhanced lifetime responsivity and tripled detection SNR and d′ over the small-amplitude (S) spikes. L spikes occurred independent of subthreshold waveform level but synchronized with spiking on plateaus shown in the inset in (F). (G) and (H) show average spike waveforms for color-coded regions. The point of initiation for S spikes was a central region of the axon (consistent with movies S1 and S2), whereas L spikes were diffuse and associated with background fluorescence. L spikes also correspond to local spikes in the dendrite and soma in (H). The L-spike background component possibly resulted from out-of-focus neurons (movies S8 and S9). In (I) and (J), histograms of action potential amplitudes are compared. In intensity, the L and S populations were not resolved and strongly overlap, but they were clearly separated in lifetime. Lifetime is used to identify the spikes, and the intensity histogram in (I) is shaded with two colors to show overlapping populations. (K) A polar histogram demonstrates strong phase locking of the L spikes to mechanical stimulus by referencing phase to a bandpass-filtered lifetime trace. S spikes do not show phase locking. icos qið Þ=NAP, is plotted versus bandpass (L) The average phase vector length, center frequency to show narrow-band locking response. P Bowman et al., Science 380, 1270–1275 (2023) 23 June 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E recordings (figs. S5 and S6). When we then applied a 10-frame moving average, the spatial distribution of slower subthreshold voltage signals could also be studied. Subthreshold signals were often strongest and localized in the dendrites (movies S6 and S7). Still frames from these movies are shown in Fig. 1I and figs. S7 and S8. By measuring a ratio of simultaneous in- tensities, EO-FLIM removes noise sources that are common mode to both the modulated channels. In Fig. 2, we show example record- ings of MBON-g1pedc>a=b neurons from six flies, comparing intensity DF =F with lifetime readouts. In all recordings, lifetime detec- tion enhanced the SNR for action potential detection by approximately twofold. This SNR was quantified by comparing spike ampli- tude with the high-frequency noise floor. We also analyzed traces by using the spike de- tection fidelity, d′, a discriminability index that quantifies spike detection by comparing the statistical distributions of spike ampli- tudes and background-noise fluctuations (21). Lifetime detection improved d′ by 1.5 to 2.4 times (Fig. 2, A to E). Noise power was also compared across tem- poral frequencies, demonstrating broad sup- pression of intensity noise (Fig. 2, A to E) and showing that EO-FLIM approaches the pho- ton shot-noise limit. To allow a direct compar- ison of noise power spectrum, the responsivity of spikes in intensity and lifetime channels was normalized. For flies not displaying much motion (Fig. 2, A and B), lifetime primarily reduced technical noise at high frequencies that resulted from the excitation laser (4 to 7 dB). Even in these well-behaved examples, lifetime readout resulted in improved SNR and d′. Lifetime recordings also improved long-term stability in the voltage readout. Intensity-based voltage imaging often displays strong motion artifacts, which degrade stability. In Drosophila, these artifacts result from movements such as extension of the proboscis. For flies displaying motion (Fig. 2, C to E, and fig. S9), lifetime readout suppressed artifacts at low frequen- cies by up to 9 dB as compared with intensity readout (in these examples, subthreshold wave- forms may only be resolved in lifetime). Figure 3 shows trace stability quantified according to the spike-amplitude distribution and the uni- formity of threshold voltage level at action potential locations. Histograms of mean nor- malized spike heights and mean normalized subthreshold level are plotted for each spike, showing that lifetime improves spike unifor- mity by up to 2.5 times and threshold unifor- mity by up to 5.8 times. With the improved readout stability af- forded by lifetime recording, we observed two spike amplitudes (Fig. 2, F to L). The small-amplitude (S) spikes occurred on top Fig. 3. Lifetime improves uniformity in action potential amplitude and threshold level. (A to D). Histograms of action potential amplitudes (lifetime in blue) and action potential levels on the subthreshold waveform (lifetime in green) are plotted for each activity trace, overlayed on the same histograms for intensity (gray). Action potential amplitudes are normalized to the mean. Subthreshold level is also mean normalized as (cid:3) (cid:2) =(cid:1) L (cid:2) (cid:1) L, where L is the spike’s corresponding level on a low-pass filtered trace, and its distance is L (cid:1) measured relative to the mean level of all other spikes, L. A perfectly uniform threshold would thus result in L = 0 for all spikes. In each histogram, the ratio of standard deviation between intensity and lifetime readouts, sF=st, is given as a figure of merit for uniformity. (A) and (B) to (D) correspond to Fig. 2, A and C to E, respectively. of subthreshold voltage plateaus, whereas large-amplitude (L) spikes were observed in bursts that were independent of subthresh- old voltage level. In this sample, the fly was mechanically stimulated with sound waves near a resonance of the microscope stage. Figure 2, I and J, show histograms of spike heights in both intensity and lifetime. The two spike populations were clearly distin- guished by using lifetime readout but were not separable with intensity readout. The L spikes showed enhanced lifetime responsiv- ity (1.34 DF =F, compared with 0.68 DF =F for S spikes). The difference in responsivity may indicate kinetic differences in the GEVI re- sponse to different action potential waveforms (3). Spike-triggered averages (movies S8 and S9) showed that the L spikes originated dif- fusely across the image, whereas the S spikes initiated in a local area of the axon, as shown in movies S1 and S2. The L spikes may be as- sociated with out-of-focus neurons displaying off-target GEVI expression and were also syn- chronous with spiking of the targeted neuron (Fig. 2, F to H). The L spikes displayed strong phase locking in response to mechanical stim- ulus, whereas the S spikes did not. A histogram of phases is displayed in Fig. 2K, where phase is determined by each spike’s location on a trace that has been filtered at the stimulus frequency. Narrow-band locking response is demonstrated in Fig. 2L. Similar L spikes were also observed during the mechanical stimulus sweep (Fig. 4C). To further demonstrate the noise-rejection capabilities of lifetime detection, we placed the fly on a piezoelectric stage to provide mechanical shaking in the xy plane. This Bowman et al., Science 380, 1270–1275 (2023) 23 June 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Phase locking of spikes to direct mechanical stimulus. (A) Lifetime provided strong rejection of intensity noise associated with shaking the sample (40-Hz square wave, ~0.8 mm peak-to-peak). (B) Intensity noise at the stimulus frequency and second harmonic were attenuated by 21 and 20 dB. (C) A spectrogram of the lifetime trace is plotted as stimulus is swept from 50 to 150 Hz. Vertical lines of activity in the spectrogram correspond to spike bursts in the lifetime trace. Mechanical cross-talk is seen as the diagonal line sweep, and phase locking appears as increased frequency content at the stimulus frequency during spike bursts. (D) To show phase locking visually, a sliding window autocorrelation of the lifetime trace is plotted with a 150-ms window. Phase locking may be seen by observing alignment of autocorrelation peaks during activity bursts to the peaks resulting from mechanical cross-talk signal. Examples of bursts showing phase locking are highlighted (blue vertical bars). direct mechanical stimulus resulted in high levels of intensity noise that obscured neu- ron activity, but lifetime readout suppressed this noise by up to 21 dB (Fig. 4, A and B). Using this direct stimulus, we could observe phase-locked spiking behavior from 30 to >100 Hz (Fig. 4 and fig. S10). As shown in these figures, we observed phase locking through increased spectral power at the stim- ulus frequency in a spectral waterfall plot as the excitation frequency was swept. This phase locking may also be visualized with the autocorrelation of the lifetime trace (Fig. 4, C and D). These observations are consis- tent with previous studies on mechanical and auditory effects in Drosophila that iden- tified a broad auditory response across the central brain (22) and responses to substrate vibrations (23, 24). Discussion EO-FLIM may be applied to both existing lifetime-sensitive probes and to donor read- out of FRET-based biosensors. Standard FRET sensors are read out by the ratio of optical intensities in spectrally separated donor and acceptor channels (25), requiring an acceptor molecule with high quantum yield. Two-color readout frequently limits detection sensitivity because of spectral cross-talk and also pre- vents probe multiplexing (26). Lifetime mea- surement removes these limitations and allows quantitative FRET measurements with only the donor channel. In voltage imaging, lifetime will enable improved measurement of FRET-opsin (3, 7), hybrid FRET (6), and dye indicators (8). We also anticipate application to imaging cal- cium (27), neurotransmitters (28, 29), and cyclic adenosine monophosphate (30). Use of GEVIs in vivo is often accompanied by a large fluorescence background that re- sults from protein expression outside the cel- lular membrane (31) or leaky gene expression from nontargeted cells (32). This background signal had a different fluorescence lifetime and photobleaches at a different rate, resulting in slow drifts of the measured lifetime traces (fig. S11). (We expect that in vitro studies will not be affected by such backgrounds.) To mea- sure absolute voltage, signal and background populations would need to be unmixed by discriminating between two closely spaced ex- ponential decays. For this reason, we focus on the improved stability and noise performance afforded by lifetime measurement rather than on absolute quantification. In our current im- plementation, we measured the population- weighted average lifetime by acquiring images at a single modulation phase. In the future, multiple modulation phases can be combined to unmix lifetime components and improve absolute measurement. We have shown that EO-FLIM enables flu- orescence lifetime imaging of neuron activity in vivo, overcoming the throughput and sensi- tivity limitations of existing FLIM techniques. We expect straightforward application to other systems, including mammalian brains, which feature both larger neurons and action poten- tials as compared with those of Drosophila (3, 7). Voltage imaging in neuroscience is one example application, but membrane potential is also broadly interesting throughout biology, from bacteria (33, 34) and plants (35) to car- diac (36–38) and muscle tissue (39). The ability of EO-FLIM to strongly reject motion noise in vivo is relevant for brain- and cardiac-imaging applications in which it is challenging to faith- fully distinguish voltage dynamics from mo- tion and hemodynamic artifacts (15, 16, 36, 37). It may become possible to perform voltage imaging during natural movements such as insect flight, or while tracking a freely moving organism (40). Further, recent advances in GEVI probes have enabled voltage imaging of populations of neurons (3, 5). Lifetime imag- ing will establish accurate long-term readout of subthreshold activity across a neural cir- cuit, allowing functional connectivity mapping where spike activity may be correlated with subthreshold modulation of downstream neu- rons. Last, by using lifetime detection in com- bination with optogenetic tools (4, 41), it will be possible to improve techniques for targeted optical activation and control (42) of neuron membrane potential. Bowman et al., Science 380, 1270–1275 (2023) 23 June 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E RE FE RENCES AND N OT ES 1. T. H. Kim, M. J. Schnitzer, Cell 185, 9–41 (2022). 2. M. B. Ahrens, M. B. Orger, D. N. Robson, J. M. Li, P. J. Keller, Nat. Methods 10, 413–420 (2013). 3. M. Kannan et al., Science 378, eabm8797 (2022). 4. H. Tian et al., All-optical electrophysiology with improved genetically encoded voltage indicators reveals interneuron network dynamics in vivo, bioRxiv 2021.11.22.469481 [Preprint] (2021). https://doi.org/10.1101/2021.11.22.469481. 5. K. D. Piatkevich et al., Nature 574, 413–417 (2019). 6. A. S. Abdelfattah et al., Science 365, 699–704 (2019). 7. Y. Gong et al., Science 350, 1361–1366 (2015). 8. P. Liu, E. W. Miller, Acc. Chem. Res. 53, 11–19 (2020). 9. R. Datta, T. M. Heaster, J. T. Sharick, A. A. Gillette, M. C. Skala, J. Biomed. Opt. 25, 071203 (2020). 10. D. Brinks, A. J. Klein, A. E. Cohen, Biophys. J. 109, 914–921 (2015). 11. J. R. Lazzari-Dean, A. M. Gest, E. W. Miller, eLife 8, e44522 (2019). 12. M. Raspe et al., Nat. Methods 13, 501–504 (2016). 13. A. J. Bowman, M. A. Kasevich, ACS Nano 15, 16043–16054 (2021). 14. A. J. Bowman, B. B. Klopfer, T. Juffmann, M. A. Kasevich, Nat. Commun. 10, 4561 (2019). 15. J. D. Marshall et al., Cell 167, 1650–1662.e15 (2016). 16. M. S. Creamer, K. S. Chen, A. M. Leifer, J. W. Pillow, PLOS Comput. Biol. 18, e1010421 (2022). 17. S. Sinha et al., Proc. Natl. Acad. Sci. U.S.A. 110, 18374–18379 (2013). 18. C. Huang et al., Nat. Commun. 9, 872 (2018). 19. R. M. Clegg, in FRET and FLIM Techniques, T. W. J. Gadella, Ed., vol. 33 of Laboratory Techniques in Biochemistry and Molecular Biology (2009), pp. 1–57. 20. C. Huang et al., Dopamine signals integrate innate and learnt valences to regulate memory dynamics, Research Square [Preprint], Version 1 (2022). https://assets.researchsquare.com/ files/rs-1915648/v1_covered.pdf?c=1660253275. 21. B. A. Wilt, J. E. Fitzgerald, M. J. Schnitzer, Biophys. J. 104, 51–62 (2013). 22. D. A. Pacheco, S. Y. Thiberge, E. Pnevmatikakis, M. Murthy, Nat. Neurosci. 24, 93–104 (2021). sub-threshold activity in vivo, Dryad (2023); https://doi.org/ 10.5061/dryad.nzs7h44wd. 23. P. Hehlert, W. Zhang, M. C. Göpfert, Trends Neurosci. 44, 44. A. J. Bowman, C. Huang, M. J. Schnitzer, M. A. Kasevich, 323–335 (2021). 24. W. Zhang, Z. Yan, L. Y. Jan, Y. N. Jan, Proc. Natl. Acad. Sci. U.S.A. 110, 13612–13617 (2013). 25. W. Akemann et al., J. Neurophysiol. 108, 2323–2337 (2012). 26. D. M. Grant et al., Biophys. J. 95, L69–L71 (2008). 27. F. H. van der Linden et al., Nat. Commun. 12, 7159 (2021). 28. P. Ma et al., Fluorescence lifetime enables high-resolution analysis of neuromodulator dynamics across time and animals, bioRxiv 2022.09.28.510014 [Preprint] (2022). https://doi.org/ 10.1101/2022.09.28.510014. 29. L. Sistemich, P. Galonska, J. Stegemann, S. Kruss, Angew. Chem. Int. Ed. Engl. 62, e202300682 (2023). 30. J. Klarenbeek, J. Goedhart, A. van Batenburg, D. Groenewald, K. Jalink, PLOS ONE 10, e0122513 (2015). 31. J. Platisa, V. A. Pieribone, Curr. Opin. Neurobiol. 50, 146–153 (2018). 32. Y. Aso et al., eLife 3, e04577 (2014). 33. A. Prindle et al., Nature 527, 59–63 (2015). 34. J. M. Kralj, D. R. Hochbaum, A. D. Douglass, A. E. Cohen, Science 333, 345–348 (2011). 35. N. B. C. Serre et al., Nat. Plants 7, 1229–1238 (2021). 36. C. D. Acker, P. Yan, L. M. Loew, Prog. Biophys. Mol. Biol. 154, 3–10 (2020). 37. P. Lee et al., Cardiovasc. Res. 115, 1659–1671 (2019). 38. J. P. Wikswo Jr., S. F. Lin, R. A. Abbas, Biophys. J. 69, 2195–2210 (1995). 39. N. Azimi Hashemi et al., Proc. Natl. Acad. Sci. U.S.A. 116, 17051–17060 (2019). 40. J. P. Nguyen et al., Proc. Natl. Acad. Sci. U.S.A. 113, E1074–E1081 (2016). 41. Y. Adam et al., Nature 569, 413–417 (2019). 42. A. C. F. Bergs et al., Nat. Comm. 14, 1939 (2023). 43. A. J. Bowman, C. Huang, M. J. Schnitzer, M. A. Kasevich, Wide-field fluorescence lifetime imaging of neuron spiking and Wide-field fluorescence lifetime imaging of neuron spiking and sub-threshold activity in vivo, Version 1, Zenodo, (2023); https://doi.org/10.5281/zenodo.7706488. AC KNOWLED GME NTS Funding: We acknowledge funding from the Gordon and Betty Moore Foundation; the US Department of Energy, Office of Science, Office of Biological and Environmental Research, under award DE-SC0021976; NIH grant U01NS120822 (M.J.S. and G. Vasan); and NSF NeuroNex grant DBI-1707261 (M.J.S. and K. Deisseroth). A.J.B. acknowledges support from the NSF Graduate Research Fellowship under grant 1656518 and the Stanford Graduate Fellowship. Author contributions: A.J.B. developed the microscope. C.H. prepared Drosophila for imaging. A.J.B. and C.H. performed the experiments. A.J.B. and M.A.K. analyzed data and wrote the manuscript. All authors contributed to experiment conception and manuscript revision. Competing interests: A.J.B. and M.A.K. are inventors on PCT/US2019/062640, US17/153438, and US17/ 898093. Data and materials availability: All data are available in the manuscript, in the supplementary material, or deposited at Dryad (43). Code is available at Zenodo (44). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adf9725 Materials and Methods Figs. S1 to S11 Movies S1 to S9 MDAR Reproducibility Checklist View/request a protocol for this paper from Bio-protocol. Submitted 23 November 2022; accepted 16 May 2023 10.1126/science.adf9725 Bowman et al., Science 380, 1270–1275 (2023) 23 June 2023 6 of 6
10.1126_science.adi1826
RES EARCH FISHERIES Large females connect Atlantic cod spawning sites Esben Moland Olsen1,2*, Ørjan Karlsen3, Jon Egil Skjæraasen3 The Earth’s ecosystems are increasingly deprived of large animals. Global simulations suggest that this downsizing of nature has serious consequences for biosphere functioning. However, the historical loss of large animals means that it is now often impossible to secure empirical data revealing their true ecological importance. We tracked 465 mature Atlantic cod (Gadus morhua) during their winter spawning season and show that large females (up to 114 centimeters in length), which are still found in mid-Norway, were characterized by more complex movement networks compared with smaller females. Large males were sparse but displayed similar movement patterns. Our finding implies that management programs promoting large fish will have positive impacts on population resilience by facilitating the continued use of a diversity of spawning habitats and the connectivity between them. C urrent ecosystems are often character- ized by an absence of large animals. This downsizing of nature is part of the on- going biodiversity crisis and involves the loss of larger species as well as the larger individuals within a species, both of which can be linked to historic and ongoing harvesting practices by humans (1). Such signatures of human activity can be traced back for millen- nia (1–3), which makes it difficult to infer the potential role of larger animals in intact eco- systems. Model simulations suggest that larger animals could have major influences on pop- ulation dynamics, ecosystem functioning, and resilience to environmental change (4), and therefore they should be the focus of biodiver- sity restoration programs (5). For instance, large predators may impose strong top-down control and prevent destabilizing grazer out- breaks (6) and even promote genetic diversity at lower trophic levels (7). Large animals could also play a particularly important role in con- necting ecosystems (8). Here, we describe the role of large female Atlantic cod (Gadus morhua), which are com- parable in body size to those typically reported in archaeological studies of cod bones dating back at least one millennium (2, 9), in con- necting coastal spawning sites. The Atlantic cod is a potentially dominant predator in North Atlantic coastal ecosystems, and indi- viduals can grow to reach a body length well beyond 1 m (2, 10). However, cod is also a prized catch in fisheries, and many populations have been seriously overfished, in some cases col- lapsed to a state where recovery is expected to be slow or even unlikely (11). Size-selective and intense fishing has also changed the life history of this fish and caused major trun- cations of historic age and size distributions 1Institute of Marine Research; Flødevigen, Arendal 4817, Norway. 2Centre for Coastal Research, Department of Natural Sciences, University of Agder; Kristiansand 4604, Norway. 3Institute of Marine Research; Bergen 5817, Norway. *Corresponding author. Email: esbenmo@imr.no (9, 12). Larger female cod are highly fecund and may spawn >10 million eggs during a sea- son (13, 14). Larger females also typically pro- duce higher-quality eggs and could therefore play an important role in population replen- ishment compared with the smaller conspe- cifics typically observed in today’s populations (14–17). Females are thought to visit male ter- ritories for pairing and spawning (18, 19), and multiyear homing to specific spawning grounds is known (20–22). One study also detected in- dividual movements among spawning sites (23). However, the environmental and pheno- typic predictors of such behavior remain un- clear. The potential for reproduction to happen at multiple locations clearly exists because fe- male cod are batch spawners. Typically, portions of eggs are matured and spawned approximate- ly every 2 to 4 days throughout the spawning season (15), with the number of batches increas- ing with maternal body size (16). We used acoustic telemetry (24) to map how individual female cod move during the spawn- ing season within a network of coastal spawn- ing sites in mid-Norway (Fig. 1). The spawning sites were identified as traditional fishing grounds where cod will aggregate during spawn- ing and by scientific surveys on cod egg and juvenile distributions (25). At the onset of three consecutive spawning seasons (2017 to 2019), a total of 213 mature female cod were captured, tagged, and tracked within these spawning habitats (table S1) (26). The female cod ranged in body size from 42 to 114 cm and from 0.8 to 15.1 kg. A total of 24% (n = 50) of the tracked females were 80 cm or larger (27). Effects of cod body size on connectivity Whereas the smaller females were typically de- tected at a few neighboring sites only, the larger females were more mobile and were detected at more, and sometimes distant, sites (Fig. 1). These basic observations were confirmed by statistical analyses using a network approach in which the acoustic receivers deployed at fixed sites represent the nodes, and fish move- ments between receiver sites represent the links (26–28). In our study, each fish was as- signed a connectivity score representing the number of specific links detected during the spawning season (29). A residency index (RI), defined as the number of days that a particular cod was detected within the telemetry array during the spawning season (30), was included as a covariate in the statistical models. This was to account for the fact that some individuals are likely to be less strongly associated with the chosen study area or moving outside the listening range of the receivers for parts of the study duration. In our case, 21% of the females (n = 45) were detected for ≥75% of the days during the spawning season, whereas 49% (n = 104) were detected for <25% of the spawning season days. After an initial filtering (26), data on all females across observed sizes and resi- dencies were included in the same statistical model selection (table S2 and fig. S2). The most parsimonious model for predicting a connectiv- ity score included an effect of body size, whereas alternative models excluding the effect of body size received very little support and were thus not considered for inference (table S2). A posi- tive effect of body size on movement among sites was very clear for resident individuals having a stronger overall association to the area covered by the telemetry array, and it was less clear for individuals only detected for mi- nor parts of the spawning period (Fig. 2). The model predicted a connectivity score of 3.3 links [95% confidence interval (CI) = 2.6 to 4.2] for resident females (RI = 0.75) with a body size of 50 cm, compared with a predicted connectivity score of 8.2 links (95% CI = 5.9 to 11.3) for large (100 cm) resident females (table S3). A 100% increase in female cod body size was thus as- sociated with a 148% increase in connectivity score. There was also a moderately negative correlation between female cod body size and residency (Pearson correlation coefficient = –0.20, P = 0.004), indicating that large females tended to move beyond the spawning sites in- cluded in our study area. Note that this cor- relation between the two predictor variables did not introduce a serious issue of collinearity when interpreting model predictions (variance inflation factor = 1.04). In a second step, we analyzed the tracked movements of the 252 male cod included in the study (table S1 and fig. S2). These males ranged in body size from 42 to 110 cm, although only 5% (n = 12) were 80 cm or larger. A total of 30% (n = 75) of the males had an RI of ≥0.75, whereas 38% (n = 95) had an RI <0.25. As for females, the larger males tended to have more extensive movement networks compared with smaller males (Fig. 1). For small and resident males (length = 50 cm, RI = 0.75), the model predicted a connectivity score of 3.4 links (95% CI = 2.8 to 4.1; Fig. 2). By comparison, for large and resident males (length = 100 cm, Olsen et al., Science 382, 1181–1184 (2023) 8 December 2023 1 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Connectivity among Atlantic cod spawning sites. (A) Individual cod were tracked using acoustic telemetry in the Smøla region of northwestern Norway (red dot). (B to E) Red lines represent the combined movement networks (links) contributed by the 10 largest females [(B), 91 to 114 cm] versus the 10 smallest females [(C), 42 to 54 cm], the 10 largest males [(D), 79 to 110 cm], and the 10 smallest males [(E), 42 to 50 cm] present during at least 20% of the spawning season, inferred from detections stored by moored acoustic receivers (blue dots). RI = 0.75), the predicted connectivity score was 7.1 links (95% CI = 5.5 to 9.3; Fig. 2 and table S3). In contrast to the females, there was no clear evidence that larger males were less resident than smaller males (Pearson correla- tion coefficient = –0.03, P = 0.61). Cod tracked in this study were captured and released at five different locations within the study area (26). In a separate analysis, we ex- amined the probability of cod moving among these more distant locations. In contrast to the finer-scale network analyses described above, each of the release locations refers to a larger area associated with multiple acoustic receiv- ers (fig. S1). A total of 38% (n = 80) of the tagged female cod were detected at more than one release location during the spawning season. The probability of being detected at multiple release locations increased significantly with body length, estimated at 0.26 (SE = 0.06) for a 50-cm female compared with 0.56 (SE = 0.09) for a 100-cm female (table S4). A total of 27% (n = 68) of the tagged male cod were detected at multiple release locations. The probability of being detected at multiple release locations increased significantly with body length, esti- mated at 0.11 (SE = 0.03) for a 50-cm male compared with 0.83 (SE = 0.09) for a 100-cm male (table S4). Connectivity and productivity of populations Our study suggests that the contribution of large female fish to population productivity and stability may go beyond what is already recognized from their high-reproductive-energy output (17, 31). From a population perspective, a network of spawning sites can indeed act to stabilize overall recruitment of broadcast spawning fish through a connectivity portfolio effect in which some sites are successful in some years, whereas other sites are successful in other years (32). Similarly, for Atlantic cod, the exact location where the pelagic eggs are spawned will influence in which habitats the juveniles eventually settle for growth and sur- vival (33, 34). There is evidence for spatial asynchrony in recruitment and juvenile growth of cod occurring on a scale of only a few tens of kilometers, including our study region (25, 35), suggesting that what stands out as the most favorable spawning site varies temporally. From the perspective of individual fitness and adaptive evolution, batch spawning by cod func- tions as part of a bet-hedging strategy that is advantageous in unpredictable environments such as the ocean (31). Recent modeling studies suggest that female cod can increase their fit- ness by spreading the risk across several re- productive events through a spawning season (36, 37). Our study shows that this bet-hedging strategy likely involves spreading the risk in space as well as over time. Potential costs asso- ciated with such strategies might include reduced offspring quality in the batches spawned toward the end of the season (16). Furthermore, maxi- mum body size can relate to life history traits such as growth, age and size at maturation, and longevity (38–40). Natural- and human-induced mortality and selection acting on these life his- tory traits may therefore also influence the bet-hedging strategy and level of connectivity displayed by female cod. Linked to life his- tories, age and level of experience could also play a role in determining which spawning sites are visited and used by cod. There are many examples of social learning among fish (41), and it has been suggested that smaller cod may follow the lead of larger conspecifics on their migrations (42). Although age and size of cod are correlated, we do not know the exact age of the fish tracked in our study. The reason is that cod age is usually determined from seasonal growth zones in their otoliths (ear bones), and this technique requires that fish are euthanized. Increased spatial bet hedging (i.e., spawn- ing at multiple sites during a season) among larger females is, in our opinion, a parsimonious Olsen et al., Science 382, 1181–1184 (2023) 8 December 2023 2 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. Connectivity response to cod body size. Shown are Atlantic cod movement links as predicted for resident females [(A), RI = 0.75, detected 75% of spawning season], females with low residency [(B), RI = 0.25], resident males [(C), RI = 0.75], and males with low residency [(D), RI = 0.25] showing the mean prediction (black lines) and 95% CI (gray shading) within the range of observed body lengths. The structure of the underlying statistical models and associated parameter estimates are provided in tables S2 and S3. interpretation of the observation that larger females tended to have a higher connectiv- ity score. In addition, other mechanisms might also be playing out. Earlier studies on spawn- ing cod in captivity have revealed a complex mating behavior in which male courtship displays and vocalizations precede pairing and spawning during characteristic “ventral mounts” (18, 43). These behaviors potentially allow females to assess male quality and to ex- hibit nonrandom mate choice (43). Therefore, it is possible that larger females were more active in seeking out and assessing potential mates while not necessarily spawning dur- ing all visits to all locations. Their behavior would nevertheless add to the spatial complex- ity of the mating system. The mating system of the cod has been de- scribed as a lek, in which at least some domi- nant males form clustered mating territories and engage in aggressive interactions toward other males, and male dominance is positively associated with body size (18, 43). Our finding that male movements among sites was com- parable to that of females suggests that the exact location of their territory is not fixed, but rather may shift during a spawning season, and that larger males are exploring more sites. Similarly, a study from the Gulf of Maine con- cluded that male cod probably do not remain faithful to a specific site when they form tem- porary individual territories (19). Our study shows that downsizing of cod by fisheries, a widespread phenomenon, may con- tribute to substantially reduced connectivity and a fragmentation of spawning habitat use, which is likely to have a negative impact on the overall population productivity (44). At a larger spatial scale and population level, catch data from fisheries that were collected across one century suggest that depletion of old-growth age structure is associated with a long-term truncation of the spawning migration of North- east Arctic cod (45) and a concurrent decline in recruitment success (46). Fisheries are pre- dicted to disarm cod risk-spreading strategies by selecting against the largest individuals (37) and perhaps also by selecting directly against exploratory behaviors (47, 48). Fishing may also effectively disrupt cod spawning aggre- gations (49). Furthermore, populations of At- lantic cod that are severely depleted by fisheries provide evidence for Allee effects in which the per capita population growth rate declines at low population densities, which will impair the ability of the population to recover should fishing pressure be reduced (11). Allee effects could be caused by mating systems becoming increasingly dysfunctional as population den- sity decreases (50). In particular, cod spawning aggregations tend to be male dominated (51), so connectivity by large females may have an important influence on the operational sex ratio and therefore the functioning of the mating sys- tem at a given point in time. We acknowledge that the importance of large individuals for population productivity, and thus the appropriate management and conser- vation measures, is still a matter of debate. For instance, Andersen et al. (52) pointed out that the largest female fish will usually be rare, so by accounting for typical fish demography, their expected impact on population replen- ishment will be limited. Similarly, a recent field study linking parents and offspring of a long-lived reef fish found that whereas re- productive success increased markedly with maternal body size, the numerous small ma- ture females were responsible for a relatively large proportion of offspring replenishment even though the parents were sampled from protected sites within marine reserves (53). The latter study therefore advocated traditional minimum size limits as a useful management tool to complement the benefits offered by ma- rine reserves. By contrast, another recent study concluded that for 32 of the world’s largest fisheries, the typical assumption that egg pro- duction is simply proportional to spawner bio- mass will underestimate the contribution of larger individuals and, consequently, substan- tially overestimate the population reproduc- tive potential (54). For northern cod found off Newfoundland and Labrador, the egg contri- bution from older females (≥10 years of age) has declined from an estimated 30 to 40% in the 1960s to a negligible level in more recent postcollapse years (31). To restore a safe op- erating space (55) for fisheries, strategies for maintaining the full potential of fish spawn- ing habitat use, set by ecological and evolu- tionary constraints (56), should be integrated into management and conservation plans. To that end, a widely supported recommenda- tion for rebuilding size and age structures is to implement slot-size limits in fisheries and fully protected marine reserves connected by seascape movement corridors (44, 54, 57, 58). Conclusions The mobility expressed by large female cod during the spawning season shows that these fish can play an important role in facilitating connectivity among spawning habitats. Man- agement actions directed toward rebuilding fish life histories could therefore be a mean- ingful way of improving overall population re- silience. Even though the populations of cod in mid- and northern Norway are harvested and far from pristine (59), the continued ex- istence of some very large individuals presents a valuable glimpse of what fully expressed and naturally adapted life histories may add in terms of movement behavior diversity. Unless the genetic basis for such life histories and behaviors has also been eroded by fisheries (48, 60), our finding offers a positive perspec- tive for conservation programs aimed at re- building more severely depleted populations and species. REFERENCES AND NOTES 1. F. A. Smith, R. E. Elliott Smith, S. K. Lyons, J. L. Payne, Science 360, 310–313 (2018). J. B. C. Jackson et al., Science 293, 629–637 (2001). 2. 3. M. Sodeland et al., Proc. Natl. Acad. Sci. U.S.A. 119, e2114904119 (2022). 4. B. J. Enquist, A. J. Abraham, M. B. J. Harfoot, Y. Malhi, C. E. Doughty, Nat. Commun. 11, 699 (2020). 5. A. H. Schweiger, J.-C. Svenning, People Nat. 2, 29–41 (2020). 6. R. L. Selden, S. D. Gaines, S. L. Hamilton, R. R. Warner, Proc. Biol. Sci. 284, 20161936 (2017). 7. E. Foster et al., Science 374, 333–336 (2021). 8. J. J. Williams, Y. P. Papastamatiou, J. E. Caselle, D. Bradley, D. M. P. Jacoby, Proc. Biol. Sci. 285, 20172456 (2018). J. H. Barrett, J. Fish Biol. 94, 1033–1044 (2019). 9. 10. K. T. Frank, B. Petrie, J. S. Choi, W. C. Leggett, Science 308, 1621–1623 (2005). 11. J. A. Hutchings, Proc. Biol. Sci. 282, 20150654 (2015). 12. E. M. Olsen et al., Nature 428, 932–935 (2004). 13. E. Oosthuizen, N. Daan, Neth. J. Sea Res. 8, 378–397 (1974). 14. G. Marteinsdottir, G. A. Begg, Mar. Ecol. Prog. Ser. 235, 235–256 (2002). Olsen et al., Science 382, 1181–1184 (2023) 8 December 2023 3 of 4 RES EARCH | R E S E A R C H A R T I C L E 15. O. S. Kjesbu, J. Fish Biol. 34, 195–206 (1989). 16. N. E. Roney, R. A. Oomen, H. Knutsen, E. M. Olsen, J. A. Hutchings, ICES J. Mar. Sci. 75, 1353–1361 (2018). 17. D. R. Barneche, D. R. Robertson, C. R. White, D. J. Marshall, 36. S. Hočevar, J. A. Hutchings, A. Kuparinen, Evol. Appl. 14, 58. K. Grorud-Colvert et al., Science 373, eabf0861 1980–1992 (2021). (2021). 37. S. Hočevar, J. A. Hutchings, A. Kuparinen, Proc. Biol. Sci. 289, 59. K. M. Norderhaug, K. Nedreaas, M. Huserbråten, E. Moland, 20221172 (2022). Ambio 50, 163–173 (2021). Science 360, 642–645 (2018). 38. A. Folkvord et al., Can. J. Fish. Aquat. Sci. 71, 1106–1112 60. D. O. Conover, S. B. Munch, Science 297, 94–96 18. V. M. Brawn, Behaviour 18, 177–198 (1961). 19. M. J. Dean, W. S. Hoffman, D. R. Zemeckis, M. P. Armstrong, ICES J. Mar. Sci. 71, 1474–1489 (2014). (2014). (2002). 39. J. A. Hutchings, A Primer of Life Histories: Ecology, Evolution, and Application (Oxford Univ. Press, 2021). AC KNOWLED GME NTS 20. D. Robichaud, G. A. Rose, Can. J. Fish. Aquat. Sci. 58, 40. A. Kuparinen, E. Yeung, J. A. Hutchings, Acta Oecol. 119, 2325–2329 (2001). 21. J. E. Skjæraasen, J. J. Meager, Ø. Karlsen, J. A. Hutchings, A. Fernö, ICES J. Mar. Sci. 68, 1472–1477 (2011). 22. D. R. Zemeckis, W. S. Hoffman, M. J. Dean, M. P. Armstrong, S. X. Cadrin, ICES J. Mar. Sci. 71, 1356–1365 (2014). 23. D. R. Zemeckis et al., ICES J. Mar. Sci. 74, 1780–1796 (2017). 24. N. E. Hussey et al., Science 348, 1255642 (2015). 25. K. Dunlop et al., Estuar. Coast. Shelf Sci. 279, 108111 (2022). 26. Materials and methods are available as supplementary materials. 27. E. M. Olsen, Ø. Karlsen, J. E. Skjæraasen, Data for: Large females connect Atlantic cod spawning sites, Dryad (2023); https://doi.org/10.5061/dryad.tb2rbp074. 28. D. M. P. Jacoby, E. J. Brooks, D. P. Croft, D. W. Sims, Methods Ecol. Evol. 3, 574–583 (2012). 29. P. Landi, H. O. Minoarivelo, Å. Brännström, U. Dieckmann, Popul. Ecol. 60, 319–345 (2018). 30. M. Espinoza, E. J. Lédée, C. A. Simpfendorfer, A. J. Tobin, M. R. Heupel, Ecol. Appl. 25, 2101–2118 (2015). 31. J. A. Hutchings, R. W. Rangeley, Can. J. Zool. 89, 386–400 (2011). 32. H. B. Harrison, M. Bode, D. H. Williamson, M. L. Berumen, G. P. Jones, Proc. Natl. Acad. Sci. U.S.A. 117, 25595–25600 (2020). 33. S. H. Espeland, J. Albretsen, E. M. Olsen, T. Bodvin, ICES J. Mar. Sci. 72, 2578–2589 (2015). 34. M. B. O. Huserbråten, E. Moland, J. Albretsen, Prog. Oceanogr. 167, 116–124 (2018). 103917 (2023). 41. C. Brown, K. N. Laland, Fish Fish. 4, 280–288 (2003). 42. G. A. Rose, Nature 366, 458–461 (1993). 43. J. A. Hutchings, T. D. Bishop, C. R. McGregor-Shaw, Can. J. Fish. Aquat. Sci. 56, 97–104 (1999). 44. S. A. Berkeley, M. A. Hixon, R. J. Larson, M. S. Love, Fisheries (Bethesda, Md.) 29, 23–32 (2004). 45. A. F. Opdal, C. Jørgensen, Glob. Chang. Biol. 21, 1521–1530 (2015). 46. J. Ohlberger, Ø. Langangen, L. C. Stige, Ecol. Appl. 32, e2614 (2022). 47. E. M. Olsen, M. R. Heupel, C. A. Simpfendorfer, E. Moland, Ecol. Evol. 2, 1549–1562 (2012). 48. S. Uusi-Heikkilä et al., Evol. Appl. 8, 597–620 (2015). 49. M. J. Dean, W. S. Hoffman, M. P. Armstrong, N. Am. J. Fish. Manage. 32, 124–134 (2012). 50. S. Rowe, J. A. Hutchings, Trends Ecol. Evol. 18, 567–572 (2003). 51. M. J. Morgan, E. A. Trippel, ICES J. Mar. Sci. 53, 820–826 (1996). 52. K. H. Andersen, N. S. Jacobsen, P. D. van Denderen, Can. J. Fish. Aquat. Sci. 76, 347–349 (2019). We thank K. M. E. Schrøder, D. Lien, and E. Kanestrøm for invaluable help with sampling of fish and telemetry array maintenance; Ø. Langangen for advice on the statistical analyses; O. G. Sørøy at the MOWI seafood company for help with logistics during fish tagging operations; and two anonymous reviewers for insightful suggestions and comments on the original version of the manuscript. Funding: This work was supported by the Research Council of Norway (grant 294926 to E.M.O.) and the Norwegian Seafood Research Fund (grant 901230 to J.E.S.). Author contributions: Conceptualization: E.M.O., Ø.K., J.E.S.; Funding acquisition: E.M.O., Ø.K., J.E.S.; Investigation: E.M.O., Ø.K., J.E.S.; Methodology: E.M.O., Ø.K., J.E.S.; Project administration: J.E.S.; Visualization: E.M.O.; Writing – original draft: E.M.O., Ø.K., J.E.S.; Writing – review and editing: E.M.O., Ø.K., J.E.S. Competing interests: The authors declare no competing interests. Data and materials availability: All data and code used in the analyses are publicly available at Dryad (27). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse 53. C. P. Lavin, G. P. Jones, D. H. Williamson, H. B. Harrison, SUPPLEMENTARY MATERIALS Proc. Biol. Sci. 288, 20202714 (2021). 54. D. J. Marshall, M. Bode, M. Mangel, R. Arlinghaus, E. J. Dick, Proc. Natl. Acad. Sci. U.S.A. 118, e2100695118 (2021). 55. J. Rockström et al., Nature 461, 472–475 (2009). 56. L. Ciannelli, K. Bailey, E. M. Olsen, ICES J. Mar. Sci. 72, 285–296 (2015). science.org/doi/10.1126/science.adi1826 Materials and Methods Figs. S1 to S3 Tables S1 to S4 References (61–74) MDAR Reproducibility Checklist 35. L. A. Rogers, G. O. Storvik, H. Knutsen, E. M. Olsen, N. C. Stenseth, J. Anim. Ecol. 86, 888–898 (2017). 57. D. J. Marshall, S. Gaines, R. Warner, D. R. Barneche, M. Bode, Front. Ecol. Environ. 17, 407–413 (2019). Submitted 8 April 2023; accepted 31 October 2023 10.1126/science.adi1826 Olsen et al., Science 382, 1181–1184 (2023) 8 December 2023 4 of 4
10.1126_science.adi6354
RES EARCH PHYSICS Ergodicity breaking in rapidly rotating C60 fullerenes Lee R. Liu1,2*, Dina Rosenberg1,2, P. Bryan Changala1,2†, Philip J. D. Crowley3, David J. Nesbitt1,2,4, Norman Y. Yao3, Timur V. Tscherbul5, Jun Ye1,2* Ergodicity, the central tenet of statistical mechanics, requires an isolated system to explore all available phase space constrained by energy and symmetry. Mechanisms for violating ergodicity are of interest for probing nonequilibrium matter and protecting quantum coherence in complex systems. Polyatomic molecules have long served as a platform for probing ergodicity breaking in vibrational energy transport. Here, we report the observation of rotational ergodicity breaking in an unprecedentedly large molecule, 12C60, determined from its icosahedral rovibrational fine structure. The ergodicity breaking occurs well below the vibrational ergodicity threshold and exhibits multiple transitions between ergodic and nonergodic regimes with increasing angular momentum. These peculiar dynamics result from the molecule’s distinctive combination of symmetry, size, and rigidity, highlighting its relevance to emergent phenomena in mesoscopic quantum systems. I solated systems that break ergodicity have been explored in a variety of experimental settings, including spin glasses (1), ultra- cold neutral atoms (2, 3) and ions (4), and photonic crystals (5). These systems ex- hibit ergodicity breaking of spin configura- tions and momentum or spatial distributions. By contrast, gas-phase polyatomic molecules provide opportunities to probe the ergodicity breaking of collective (rotational and vibra- tional) excitations in a finite quantum system. In this context, a topic of major interest has been the transport of energy deposited into molecular vibrations by optical pumping or collisions. Intramolecular vibrational redis- tribution (IVR) sets in once a critical thresh- old, defined by the product of vibrational coupling and the local density of vibrational states (which has a power-law scaling with vibrational energy), is exceeded (6–11). This vibrational ergodicity transition has been studied vigorously in the context of under- standing and controlling unimolecular reac- tion dynamics (12). Among polyatomic molecules, buckmin- sterfullerene (12C60) is notable for its structural rigidity and high degree of symmetry, which suppress IVR and allow for spectroscopic res- olution (13) and optical pumping (14) of indi- vidual rovibrational states—an unusual and fortuitous situation for a molecule with 174 vibrational modes. Its small rotational con- stant and stiff, cage-like structure ensure that 1JILA, National Institute of Standards and Technology and University of Colorado, Boulder, CO 80309, USA. 2Department of Physics, University of Colorado, Boulder, CO 80309, USA. 3Department of Physics, Harvard University, Cambridge, MA 02135, USA. 4Department of Chemistry, University of Colorado, Boulder, CO 80309, USA. 5Department of Physics, University of Nevada, Reno, NV 89557, USA. *Corresponding author. Email: lee.richard.liu@gmail.com (L.R.L.); ye@jila.colorado.edu (J.Y.) †Present address: Center for Astrophysics, Harvard and Smithsonian, Cambridge, MA 02138, USA. Liu et al., Science 381, 778–783 (2023) 18 August 2023 hundreds of rotational states are populated even when vibrational excitations are largely frozen out, which can be achieved with mod- est buffer gas cooling to ~120 K. Thus, a ther- mal ensemble of 12C60 can reveal extensive, state-resolved rotational perturbations span- ning hundreds of rotational quanta by elimi- nating vibrational “hot bands.” First observed and understood in atomic nuclei (15–22), rotational perturbations can arise from spherical symmetry breaking in the frame fixed to a rotating self-bound deform- able body (23), which lifts the degeneracy of different body-fixed projections of the total angular momentum vector J. Such perturba- tions, also called “tensor interactions” because of their anisotropic nature, manifest in fine- structure splitting of the total angular momen- tum (J) multiplets in rovibrational spectra and encode rich dynamics such as rotational bi- furcations (18, 24), as previously observed in tetrahedral SnH4, CD4, CF4, SiH4, and SiF4 and octahedral SF6 molecules (25–33). Nevertheless, observing icosahedral tensor interactions, first predicted for 12C60 over three decades ago (34), has remained an elusive goal, because there are far fewer examples of icosahedral mole- cules, nonspherical interactions occur only at higher orders of interactions, and icosahedral molecules are necessarily larger than lower- symmetry spherical top molecules, implying a smaller rotational constant. In this work, we observed these icosahedral tensor interaction splittings, revealing rota- tional ergodicity transitions in 12C60 at energies well below its IVR threshold (10). Specifically, as the molecule “spins up” to higher J, the dynamics of the angular momentum vector J in the molecule-fixed frame switches between ergodic and nonergodic regimes. In the non- ergodic regime, distinct vector J trajectories exist in the same energy range but remain separated by energy barriers. In the limit of high J, the tunneling between these trajecto- ries is too weak to restore ergodicity, leaving a characteristic signature in the fine-structure level statistics. This phenomenon differs from IVR in three key respects: (i) It involves the “transport” of the molecule frame orientation of vector J instead of vibrational energy; (ii) it can occur well below the IVR threshold; and (iii) it switches back and forth multiple times between ergodic (described by a 6th rank ten- sor interaction) and nonergodic (described by a 10th rank tensor interaction) regimes as the angular momentum is varied. This peculiar dynamical behavior arises from multiple avoided crossings with other vibrational states, which induce nonmonotonic variations in the molecule’s anisotropic character as J is varied. The rotational ergodicity transitions bear some similarity to those studied in asymmetric top molecules in a static electric field (35, 36) in that both concern the transport of angular momentum in the molecule frame. However, unlike in (35, 36), the rotational ergodicity transitions in 12C60 are induced by intramolec- ular rovibrational coupling in the freely rotat- ing molecule, rather than by an externally applied electric field. Effective 12C60 rovibrational Hamiltonian The rovibrational structure of C60 can be de- scribed by a field-free molecular Hamiltonian H ¼ Hscalar þ Htensor ð1Þ The scalar Hamiltonian Hscalar contains only those combinations of vector J and vibrational angular momentum ‘ that preserve their spheri- cal degeneracy (37) Hscalar ¼ n0 þ BJ2 þ DJ4 þ ⋯ (cid:2) 2BzJ (cid:3) ‘ ð2Þ where n0 is the vibrational band origin, B is the rotational constant, D is the scalar cen- trifugal distortion constant, and z is the Coriolis coupling constant. Rovibrational fine structure is encoded in the tensor Hamiltonian Htensor. For simplicity, we considered a pure rotational tensor Hamiltonian consisting of the two lowest-order “icosahe- dral invariants.” These invariants are linear combinations of spherical tensors of the same rank that transform according to the totally symmetric irreducible representation in the icosahedral point group (Ih) (38). They can be expressed (39) in the basis of spherical har- monics Y k Þ of degree k and order q, which depend explicitly on the molecular frame’s polar q and azimuthal f angles (Fig. 1A). The first two nontrivial (anisotropic) icosahedral invariants, with ranks 6 and 10, are given by q q; fð T 6½ (cid:4) q; fð Þ ¼ p ffiffiffiffi 11 Y 6 Þ 0 q; fð 5 p ffiffiffi (cid:3) 7 5 Y 6 5 q; fð þ Þ (cid:2) Y 6 Þ (cid:2)5 q; fð (cid:4) ð3Þ 1 of 6 RES EARCH | R E S E A R C H A R T I C L E A B 0.5 0 -0.5 C π/4 D π E π/4 F π 165 170 175 180 165 170 175 180 165 170 0.4 0.2 0 0 0.5 1 0 0.5 1 0 175 J 0.5 r 180 165 170 175 180 165 170 175 180 1 0 0.5 1 0 0.5 1 Fig. 1. Rotational energy surfaces and eigenvalues corresponding to icosahe- dral invariant spherical tensors. (A) Symmetries of C60. (Left to right) (1) Ball- and-stick model of C60, with the three different types of rotational symmetry axes that label stationary points on the rotational energy surface (RES). The degeneracies of the stationary points are listed in parentheses. Color and plot marker coding for each type of rotational symmetry axis are shown. (2) Body-fixed coordinates: polar q and azimuthal f angles. (3–5) View along C5, C3, and C2 rotational Þ ¼ symmetry axes. (B to F) (Top panels) RESs, defined by their radii r q; fð Þ ð Þ=2, for n ranging from 0 to p. Eigenvalues of 1 þ H 6þ10 1 þ H 6þ10 tensor nð Þ=2 ð tensor n; q; f ð Þ calculated for the fully symmetrized J ¼ 174 subspace are plotted on the surface as radial contours, colored corresponding to their dominant rotational symmetry character. (Center panels) Tensor energy defects [eigenvalues of H 6þ10 tensor nð Þ] over a range of J. The gray vertical line highlights J ¼ 174. Eigenvalues are plotted using the marker corresponding to their dominant C5, C3, or C2 character (A). Near the separatrices, the assignment is somewhat ambiguous owing to strong mixing. (Bottom panels) Distribution p rð Þ of energy gap ratios r (see text) calculated from tensor energy defect spectra aggregated over J = 0 to 400. Away from n ¼ 0; p [(C) to (E)], the finite value of p rð Þ as r → 0 is a signature of ergodicity breaking. Þ ð T 10½ (cid:4) q; fð Þ ¼ (cid:2) p p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 (cid:3) 13 (cid:3) 19 75 ffiffiffiffiffiffiffiffiffiffiffiffi 11 (cid:3) 19 25 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3 (cid:3) 11 (cid:3) 17 75 (cid:3) p Y 10 0 q; fð Þ Y 10 5 q; fð Þ (cid:2) Y 10 Þ (cid:2)5 q; fð (cid:4) (cid:4) (cid:3) 10 q; fð Y 10 þ Þ þ Y 10 (cid:2)10 q; fð ð4Þ which are combined to construct a truncated tensor Hamiltonian Þ ð H 6þ10 tensor Þ (cid:5) (cid:4) ¼ g cosnT 6½ (cid:4) þ sinnT 10½ (cid:6) ð5Þ This Hamiltonian is parameterized by an overall scaling factor g and mixing angle n such that n ¼ 0 and n ¼ p=2 correspond to pure T 6½ (cid:4) and pure T 10½ (cid:4), respectively. All operators that are incompatible with icosahedral point group symmetry, including any spherical harmonics of rank 1 to 5, 7 to 9, 11, 13, 14,... (40), vanish from the Hamiltonian. The use of full rovibrational tensor operators is unlikely to change the picture qualitatively, particularly when J (~100 to 300) is much greater than the vibrational angular momentum quantum number ‘ ¼ 1 (38). These polyhedral invariants are similar to those used to describe the crystal field splitting of electronic orbitals owing to an external lattice environment (41) or the ligand field splitting in transition-metal complexes (42). The energetic correction, or tensor energy defect, associated with orienting J in different directions in the molecule frame can be vi- sualized by the altitude of a semi-classical “rota- tional energy surface” (RES), defined at a fixed J. Various possible icosahedral RESs, defined Þ=2, by their radii r q; fð corresponding to different mixing angles n, are plotted for J ¼ 174 in the top panels of Fig. 1, B to F. Stationary points always lie on C2, C3, or ð Þ Þ ¼ 1 þ H 6þ10 ð tensor n; q; f C5 rotational symmetry axes. However, as n varies, they change in character between mini- ma, maxima, and saddle points. The RES dic- tates the dynamics of J in the molecule frame (30, 43–46), analogous to how an adiabatic potential energy surface steers the relative motions of nuclei (47). During free evolution, the trajectory of J follows an equipotential contour of the RES. In a full quantum-mechanical treatment, the perturbation H 6þ10 Þ ð leads only to discrete tensor tensor energy defects ei given by the eigen- values of H 6þ10 ð Þ in a fully symmetrized fixed-J tensor subspace. These orbits trace out the closed contours on the RES in Fig. 1. The orbits may also be obtained directly from the RES: They are the trajectories that both (i) satisfy a Bohr quantization condition (44, 48) and (ii) trans- form according to the totally symmetric irreduc- ible representation of the icosahedral point group (38). The latter condition accounts for the Liu et al., Science 381, 778–783 (2023) 18 August 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E quantum indistinguishability of each bosonic nucleus in the 12C60 isotopolog (13, 49) and is analogous to the selection of odd or even rotational states in ortho- or para-hydrogen molecules, respectively. The tensor energy de- fects are plotted for a range of J in the center panels of Fig. 1, B to F, and may be expected to appear in the rovibrational fine structure of 12C60. Resolving 12C60 rovibrational fine structure In the preceding discussions, we have focused solely on the geometric effects of icosahedral symmetry. In general, however, the measured tensor defect spectra may exhibit additional J-dependent scaling and offsets, which depend on intramolecular couplings specific to 12C60. To resolve rovibrational perturbations in 12C60, in this work, we explored the P-branch region of the 1185 cm−1 T1u (3) band, first identified as a region of potential interest in (13). Using cavity-enhanced continuous-wave (cw) spec- troscopy with a quantum cascade laser (QCL) source, we achieved a minimum absorption sensitivity amin ¼ 2:1 (cid:5) 10(cid:2)10cm(cid:2)1 Hz(cid:2)1=2, or 1000-fold better detection sensitivity per spec- tral element than in (13) (amin ¼ 2:2 (cid:5) 10(cid:2)7 cm(cid:2)1 Hz(cid:2)1=2 per comb mode). We acquired 600-MHz-wide absorption spectra by simulta- neously scanning the QCL frequency and free spectral range of the enhancement cavity across molecular absorption lines, and recording the frequency-dependent absorption. These spec- tra were stitched together by a combination of direct calibration of the QCL frequency with a Fourier transform spectrometer and com- parison to matching spectral features in the broadband, low signal-to-noise (SNR) frequency comb spectrum of (13). We obtained an absolute frequency accuracy of ∼6 MHz throughout the entire measured frequency range, limited by the resolution of the reference frequency comb spectrum. Finally, to ensure consistency of the absorption signal over multiple days of data collection, we have periodically remea- sured the molecular absorption feature at R(J ¼ 181) at n ¼ 1186:27 cm−1 and normalized all data taken around the same time to its line strength and measured cavity finesse. Assigning 12C60 rovibrational fine structure The culmination of these efforts is the infrared spectrum in Fig. 2, spanning the spectral re- gion from 1182.0 to 1184.7 cm−1. At J ≲ 70, there is a regular progression of rotational lines, similar to those in the R-branch (13). They rapidly split into intricate patterns before merging at and beyond J ≈ 300. Zooming into the region labeled B), the rotational line cen- ters could be fit to the scalar part of Eq. 1, as was done in (13). nP Jð Þ ¼ n0 þ B′ J (cid:2) 1 Þ ð (cid:2) B″J J þ 1 ð Þ Þ J þ 2z ð ð6Þ where J here refers to the ground-state total angular momentum. The scalar centrifugal dis- tortion term was not significant at our spec- tral precision and range of J . The fit yielded B″ = 0.0028 cm(cid:2)1 for the ground-state rota- ð Þ , where B ¼ B′ þ B″ tional constant, B′ ¼ B″ (cid:2) 2:876 (cid:5) 10(cid:2)7 cm−1 for the excited-state rotational constant, z ¼ (cid:2)0:37538 for the Coriolis coupling constant, and n0 ¼ 1184:85 cm−1. Equation 6 yields a progression of rotational lines with a spacing of ∼2B 1 (cid:2) z Þ=2 . The ð spectroscopic constants were underdetermined and only served to facilitate J -assignment of the peaks in a manner consistent with the R-branch assignments of (13). The extrapo- lated P-branch spectral line positions based on this scalar fit are plotted as gray vertical lines in Fig. 2, B to F, with every fifth J value labeled in red. The agreement with the mea- sured line positions is excellent in the region J < 70 , where the spectrum appears unper- turbed (Fig. 2B). To confirm our J assignment, we compared the peak absorption cross sections to the nu- clear spin weights of the ground-state rota- tional levels and found that they match well. Finally, we applied a frequency-dependent scaling factor to the raw absorption spectrum ð2J þ 1Þ(cid:2)1eB″J Jþ1 Þ=kBT . This scaling removes the effects of lab frame angular momentum degeneracy and the thermal ensemble, em- phasizing the dynamics in the molecule-fixed frame. ð The peaks were identified manually and circled in blue. Figure 2, C and D, show two representative regions, at J ∼ 90 and J ∼ 170, respectively, where peaks could still be indi- vidually resolved. The local peak density again matches the predicted nuclear spin weights (in blue), confirming that the rovibrational A B C E D F . . . . . . . . Fig. 2. Direct continuous-wave (cw) absorption spectroscopy of C60 P-branch. (A) Complete normalized cw spectrum of P-branch to expose the J dependence of the nuclear spin weights. Red highlighted regions are shown in greater detail in subsequent panels. (B to F) Enlargement of red highlighted regions in (A). (B) Enlargement of low-J region of the P-branch. This relatively unperturbed region of the P-branch is fit to a rigid-rotor ð Hamiltonian to obtain the rotational spacing 2B 1 (cid:2) z are indicated by the gray vertical grid lines, with J labeled in red. Blue numbers denote calculated nuclear spin weights, which match the measured peak intensities well. In (C) and (D), peaks are resolvable and marked with blue circles. In (E) and (F), spectral congestion prohibits identification of individual peaks. Þ. Fitted line positions Liu et al., Science 381, 778–783 (2023) 18 August 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E fine-structure splitting originates from icosahe- dral tensor interactions of Eq. 5 that lift K - degeneracy. Figure 2, E and F, show two regions where the peaks have begun to merge, and individual peaks can no longer be easily identified (J > 247). Interpreting the tensor splittings requires assigning each absorption peak to a partic- ular J. We began by assigning each peak to its nearest J value according to the scalar fit of Eq. 6. Subtracting the scalar contribution (Eq. 6) from the central frequency of each peak generated a single “period” of a Loomis- Wood-like defect plot in Fig. 3A. There remains some ambiguity in the defect assignments, as illustrated in the inset of Fig. 3A: Each defect is Þ, constrained to the line with slope 2B 1 (cid:2) z which passes through its current position. ð By carefully rearranging individual defects according to these discrete allowed “moves,” we unwrapped five distinct regions that ex- hibit continuous-looking patterns (Fig. 3B). Because of the discontinuities at J ≃ 80, 110, 160, and 220, there was still some ambigu- ity in the overall shift of the four perturbed sections labeled (i) to (iv). To remove this ambiguity, we recognized each section’s dom- inant pure-rank tensor character as follows: (i) (cid:2)T 6½ (cid:4) ; (ii) T 10½ (cid:4) ; (iii) T 6½ (cid:4) ; (iv) (cid:2)T 6½ (cid:4) . Eigenvalue spectra in Fig. 1, B, D, and F, show that the extremal eigenvalues associated with Cn rotational symmetry axes occur when J is an integer multiple of n. The J -assignment depicted in Fig. 3B satisfies this condition for all sections simultaneously. This final J- assignment was confirmed by the excellent agreement between J -resolved peak counts and calculated nuclear spin weight far from the discontinuities Fig. 3C. Rotational ergodicity transitions The J-dependent tensor energy defects imply rovibrational dynamics of 12C60. To infer these dynamics, we parameterized each of the four perturbed regions (i) to (iv) in Fig. 3B in terms of a mixed tensor (Eq. 5), J-dependent scaling b Jð Þ, and J-dependent scalar offset a Jð Þ: a Jð Þ þ b Jð Þ (cid:5) e n; J; K ð Þ ð7Þ where e n; J; K Þ are the J-dependent tensor ð splittings as plotted in the lower panels of Fig. 1, B to F. First, a Jð Þ was obtained from the observed mean defect of each section. Next, by perform- ing a point-cloud registration (50–52) to the theoretical eigenvalue spectra and the mea- sured defect plot, we assigned a best-fit mixing angle n to each section: (i) p; (ii) 0.45p; (iii) 1.9p; and (iv) p (38). Finally, b Jð Þ was obtained from a least-squares fit to a polynomial in J (38). The resulting reconstructed defect plot for re- gions (i) to (iv) is shown in Fig. 3D. The abrupt changes in mixing angles n are associated with transitions between ergodic and nonergodic rotational dynamics as the mo- lecule “spins up” to higher J. These dynamics ) 1 - m c ( t c e f e d ) 1 - m c ( t c e f e d 40 60 80 20 0.01 0 -0.01 0.01 A 0.02 0.01 0 -0.01 -0.02 20 s t n u o c 10 5 0 20 0 -0.01 125 130 135 140 100 120 140 160 180 200 220 240 260 280 B 40 C 60 60 80 100 120 140 160 180 200 220 240 260 280 nuclear spin weight peak counts C 40 60 80 100 120 140 160 180 200 220 240 260 280 ν = π 0.45π 1.9π π D 0.02 0.01 0 -0.01 -0.02 ) 1 - m c ( t c e f e d 20 40 60 80 100 120 140 160 180 200 220 240 260 280 J Fig. 3. Obtaining P-branch tensor energy defects versus J. (A) Plot of experimental energy defects from nearest-neighbor J assignment. Peaks are assigned to nearest J according to rigid rotor model (gray vertical grid lines in Fig. 2, B to F). Red defect points (J > 247) correspond to peaks in the absorption spectrum that are no longer individually resolved. Their peak centers are obtained from fitting to a cluster of Voigt lineshapes (38). (Inset) Unwrapping procedure follows a series of allowed “moves” that simultaneously translate points in vertical steps of ±2B(1− z) and horizontal steps of ±1. (B) “Unwrapped” defect plot. Four avoided crossings with varying strengths are seen at J ≃ 80, 110, 160, and 220. Defect patterns resemble eigenvalue spectra of icosahedral invariant tensors. (C) Calculated nuclear spin weights overlaid with measured peak counts. Blue highlighted sections show excellent agreement between calculated nuclear spin weights of C60 and assigned peak counts. (D) Reconstruction of perturbed sections of P-branch spectrum based on eigenvalue spectra of mixed tensor operator H 6þ10 Four perturbed portions of (B) are reproduced with four mixing angles n. ð Þ tensor (n). Liu et al., Science 381, 778–783 (2023) 18 August 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E transition from ergodic for region (i), to nonergodic for region (ii), and back again for regions (iii) and (iv). The origin of this ergodicity breaking can be understood semi-classically using the RESs in Fig. 1, B, D, and F. The dynamics of vector J are ergodic when time evolution explores the full space of symmetry-allowed states at the same energy. For region (iii), the dominant tensor character is T 6½ (cid:4). Naïvely, the existence of 12 disconnected trajectories encircling the C5 axes breaks ergodicity. However, these traj- ectories cannot be distinguished in 12C60: Ow- ing to the indistinguishability of the 12C nuclei, all 12 semi-classical trajectories correspond to a single quantum state given by their fully symmetric superposition. As such, the vector J dynamics do explore the full range of states at a given energy, and hence are ergodic. The same argument applies for regions (i) and (iv), which differ from (iii) only in the sign of tensor energy defects (Fig. 1F). By contrast, for region (ii), the dominant (cid:4). The C5 and C3 axes tensor character is T 10½ both correspond to peaks on the RES and to host trajectories over the same range of ener- gies (Fig. 3D). Trajectories encircling the C5 and C3 axes are distinguishable, and hence cor- respond to distinct quantum states. The quan- tum tunneling between these trajectories is unable to restore the ergodicity: The tunneling integral between C5 and C3 is exponentially small in J (53, 54), whereas the level spacing only scales as 1=J. The scaling of the tunnel- ing integral follows from standard Wentzel– Kramers–Brillouin (WKB) results (55), and the scaling of level spacings can be seen by compar- ing the relatively fixed bandwidth of tensor en- ergy defects (Fig. 3B) with the nuclear spin weight ∼ 2J þ 1 Þ=60. Consequently, for our measured J ð (well within the large-J limit), tunneling correc- tions are typically only perturbative. Energy-level statistics provide a simple probe of this ergodicity breaking in molecular spec- tra (56, 57). Quantum ergodicity is associated with eigenstates extended in phase space that can be strongly coupled by local perturbations, inducing energy-level repulsion. By contrast, ergodicity breaking is associated with the ex- istence of localized eigenstates, which are not strongly coupled by perturbations and whose energies are uncorrelated (58). Ergodic and nonergodic dynamics are therefore respec- tively associated with level repulsion and its absence (59). A useful diagnostic tool is the distribution p rð Þ, where r is the ratio of consec- utive level spacings ei (60, 61): si(cid:2)1 si ri ¼ min si si(cid:2)1 ð8Þ (cid:7) (cid:8) ; si ¼ eiþ1 (cid:2) ei ð9Þ In the limit of r → 0, level repulsion in an ergodic system causes p rð Þ → 0, and for a non- ergodic system p rð Þ → constant. Similar energy- level statistics have been used to analyze systems as diverse as nuclear spectra (59), ultracold atomic scattering (62), and many-body local- ization (63). Figure 4, A to C, show the energy-gap ratios computed from sections (i) to (iii), respec- tively, of the experimental defect plot (Fig. 3B). Here, sections (i) and (iii) exhibit persistent level repulsion characteristic of ergodicity, whereas section (ii) does not, indicating non- ergodicity. We aggregated the energy-gap ratios over each one of sections (i) to (iii) and their respective distributions in Fig. 4, D to F. These distributions confirm the presence of level re- pulsion in sections (i) and (iii) and its absence in section (ii). Þ The physical origin of rovibrational tensor energy defects in 12C60 can be inferred from Fig. 3B. The defects arise from rovibrational coupling between the bright P-type excited þð 1u 3ð Þ state and a background of perturb- T ing dark states. Specifically, both the disconti- nuities and excess observed peaks at specific J values are characteristic of avoided cross- ings with perturbing zero-order dark vibrational þð Þ 1u 3ð Þ from combination states. As they cross T below, rovibrational coupling lifts the degen- þð 1u 3ð Þ state, im- eracy of J multiplets in the T parting tensor character that depends on the identity of the perturbing vibrational state. þð 1u 3ð Þ state (spe- The tensor character of the T cifically the fitted n values) is stable in between Þ Þ avoided crossings, suggesting that each of these sections is dominantly affected by just one perturbing state at a time. At the avoided crossings, this assumption breaks down, as made particularly evident by the rapid change in mixing angle just before and after the avoided crossing at J ¼ 160 of Fig. 3B. There is no apparent structure to the changes in mixing angle and coupling strength induced by the different avoided crossings, suggesting that the perturbing dark states are distinct and not part of the same Coriolis manifold. Finally, using the observed local density of per- ≈ 2=cm(cid:2)1 and average mea- turbing states robs sured vibrational coupling strength (bandwidth of the avoided crossings) of bavg ≈ 2(cid:5) 10(cid:2)2 cm(cid:2)1 (38), we arrive at an IVR threshold parameter (10) T Eð Þ ¼ robsbavg ≈ 0:06 ≪ 1.Our 12C60 rotational ergodicity transitions are observed well below the IVR threshold, a conclusion sup- ported by the spectroscopically well-resolved T1u 3ð Þ band. This further highlights the fact that, although the rotational ergodicity tran- sition relies on intramolecular vibrational cou- pling, it is completely distinct from IVR. ffiffiffiffiffiffiffiffiffiffi 2p=3 p Conclusion We have measured and characterized icosahe- dral tensor rovibrational coupling in 12C60. Analysis of the spectrum of rovibrational ten- sor energy defects revealed that as the mol- ecule is spun up to higher J, there is a series of transitions in the dynamical behavior of J in the fixed body frame. Specifically, vec- tor J switches between ergodic and nonergodic behavior at particular J values, leaving a char- acteristic imprint on the defect-level statistics. These ergodicity transitions arise from dark vibrational combination states that cross the þð 1u 3ð Þ state at particular J values, transferring T þð Þ 1u 3ð Þ their anisotropic character onto the T state through rovibrational coupling. Þ Our measurements open the door to a rich hierarchy of emergent behavior in C60 isotopo- logs, accessible at ever-higher spectral resolution. The small nuclear spin-rotation interaction— for example, in 13C-substituted isotopologs of C60—can have a magnified effect owing to the small superfine splittings near RES extrema. A 1 r 0.1 0.01 1 B 1 1 C 1 i) 0.5 0.1 ii) 0.5 0.1 90 100 J 0 0 0.2 0.4 p(r) 0.01 120 140 J 0 0 0.2 0.4 p(r) 0.01 180 J 1 0.5 0 0 0.2 0.4 p(r) iii) 200 Fig. 4. Energy-level statistics in ergodic and nonergodic regimes. (A to C) (Left panels) Gap ratio r as a function of J calculated from sections (i) to (iv) of the defect spectrum (Fig. 3B). Gap ratios are only plotted at J values for which the peak counts match the calculated nuclear spin weight (Fig. 3C). (Right panels) Normalized distribution p rð Þ of gap ratios aggregated over sections (i) to (iii). Note the change from logarithmic to linear r scale between left and right panels. Sections (i) and (iii) exhibit level repulsion, a signature of ergodicity, whereas section (ii) does not, indicating nonergodicity. Liu et al., Science 381, 778–783 (2023) 18 August 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E Such “hyperfine” coupling can lead to spon- taneous symmetry breaking in a finite system (31, 64). These insights could prove useful for exploiting the exotic orientation state space of C60 for quantum information processing (65) and for investigating the quantum-to- classical transition of information spreading (66). Ultimately, spectroscopy of C60 isotopo- logs at ever-higher spectral resolution promises to uncover deeper insights into the emergent dynamics of mesoscopic quantum many-body systems. RE FE RENCES AND N OT ES 1. K. Binder, A. P. Young, Rev. Mod. Phys. 58, 801–976 (1986). 2. T. Kinoshita, T. Wenger, D. S. Weiss, Nature 440, 900–903 (2006). 3. M. Schreiber et al., Science 349, 842–845 (2015). J. Smith et al., Nat. Phys. 12, 907–911 (2016). 4. 5. M. Segev, Y. Silberberg, D. N. Christodoulides, Nat. Photonics 7, 197–204 (2013). J. D. McDonald, Annu. Rev. Phys. Chem. 30, 29–50 (1979). 6. 7. G. M. Nathanson, G. M. Mcclelland, J. Chem. Phys. 81, 629–642 (1984). 8. G. M. McClelland, G. M. Nathanson, J. H. Frederick, F. W. Farley, Excited States, E. C. Lim, K. K. Innes, eds. (Academic Press, Inc., 1988), vol. 7, pp. 83–106. 9. K. Sture, J. Nordholm, S. A. Rice, J. Chem. Phys. 61, 203–223 (1974). 10. D. E. Logan, P. G. Wolynes, J. Chem. Phys. 93, 4994–5012 (1990). 11. D. M. Leitner, Entropy (Basel) 20, 673 (2018). 12. D. J. Nesbitt, R. W. Field, J. Phys. Chem. 100, 12735–12756 (1996). 13. P. B. Changala, M. L. Weichman, K. F. Lee, M. E. Fermann, J. Ye, Science 363, 49–54 (2019). 14. L. R. Liu et al., PRX Quantum 3, 030332 (2022). 15. 16. 17. Y. R. Shimizu, J. D. Garrett, R. A. Broglia, M. Gallardo, I. M. Pavlichenkov, Sov. Phys. JETP 55, 5–12 (1982). I. M. Pavlichenkov, B. I. Zhilinskii, Ann. Phys. 184, 1–32 (1988). E. Vigezzi, Rev. Mod. Phys. 61, 131–168 (1989). I. M. Pavlichenkov, Phys. Rep. 226, 173–279 (1993). 18. 19. S. Frauendorf, Rev. Mod. Phys. 73, 463–514 (2001). 20. H. Hübel, Prog. Part. Nucl. Phys. 54, 1–69 (2005). 21. J. Kvasil, R. G. Nazmitdinov, Phys. Rev. C Nucl. Phys. 73, 014312 (2006). 22. J. Kvasil, R. G. Nazmitdinov, A. S. Sitdikov, P. Veselý, Phys. At. Nucl. 70, 1386–1391 (2007). 23. K. Fox, H. W. Galbraith, B. J. Krohn, J. D. Louck, Phys. Rev. A Gen. Phys. 15, 1363–1381 (1977). 24. G. Pierre, D. A. Sadovskii, B. I. Zhilinskii, Europhys. Lett. 10, 409–414 (1989). 25. R. S. McDowell et al., J. Mol. Spectrosc. 83, 440–450 (1980). 26. C. W. Patterson, A. S. Pine, J. Mol. Spectrosc. 96, 404–421 (1982). 28. K. C. Kim, W. B. Person, D. Seitz, B. J. Krohn, J. Mol. Spectrosc. 61. Y. Y. Atas, E. Bogomolny, O. Giraud, G. Roux, Phys. Rev. Lett. 76, 322–340 (1979). 110, 084101 (2013). 29. W. G. Harter, J. Stat. Phys. 36, 749–777 (1984). 30. W. G. Harter, C. W. Patterson, J. Math. Phys. 20, 1453–1459 62. A. Frisch et al., Nature 507, 475–479 (2014). 63. D. A. Abanin, E. Altman, I. Bloch, M. Serbyn, Rev. Mod. Phys. (1978). 31. W. G. Harter, H. P. Layer, F. R. Petersen, Opt. Lett. 4, 90–92 (1979). 32. R. S. McDowell, H. W. Galbraith, B. J. Krohn, C. D. Cantrell, E. D. Hinkley, Opt. Commun. 17, 178–183 (1976). 33. J. Bordé, C. J. Bordé, Chem. Phys. 71, 417–441 (1982). 34. W. G. Harter, D. E. Weeks, Chem. Phys. Lett. 132, 387–392 (1986). 35. M. Abd El. Rahim, R. Antoine, M. Broyer, D. Rayane, P. Dugourd, J. Phys. Chem. A 109, 8507–8514 (2005). 36. S. M. Pittman, E. J. Heller, J. Phys. Chem. A 119, 10563–10574 (2015). 37. K. T. Hecht, J. Mol. Spectrosc. 5, 355–389 (1961). 38. Materials and methods are available as supplementary materials. 39. Y. Zheng, P. C. Doerschuk, Acta Crystallogr. A 52, 221–235 (1996). 40. P. R. Bunker, P. Jensen, Mol. Phys. 97, 255–264 (1999). 41. K. R. Lea, M. J. Leask, W. P. Wolf, J. Phys. Chem. Solids 23, 1381–1405 (1962). 42. B. J. S. Griffith, L. E. Orgel, Q. Rev. Chem. Soc. 11, 381–393 (1957). 43. W. G. Harter, Phys. Rev. A Gen. Phys. 24, 192–263 (1981). 44. W. G. Harter, C. W. Patterson, J. Chem. Phys. 80, 4241–4261 (1984). 45. W. G. Harter, Comput. Phys. Rep. 8, 319–394 (1988). 46. D. A. Sadovskii, B. I. Zhilinskii, J. P. Champion, G. Pierre, J. Chem. Phys. 92, 1523–1537 (1990). 47. M. Khauss, Annu. Rev. Phys. Chem. 21, 39–46 (1970). 48. W. G. Harter, D. E. Weeks, J. Chem. Phys. 90, 4727–4743 (1989). 49. R. D. Johnson, G. Meijer, D. S. Bethune, J. Am. Chem. Soc. 112, 8983–8984 (1990). 50. Y. Chen, G. Medioni, Image Vis. Comput. 10, 145–155 (1992). 51. P. J. Besl, N. D. Mckay, IEEE Trans. Pattern Anal. Mach. Intell. 14, 239–256 (1992). 52. A. V. Segal, D. Haehnel, S. Thrun, Robot. Sci. Syst. 5, 435 (2010). 53. A. Auerbach, Interacting electrons and Quantum magnetism (Springer Science+Business Media, 1994). 54. J. L. van Hemmen, A. Sütö, Europhys. Lett. 1, 481–490 (1986). 55. L. Landau, E. Lifshitz, Quantum Mechanics (Non-Relativistic Theory), vol. 3 (Pergamon Press, 1958). 56. M. V. Berry, M. Tabor, Proc. R. Soc. London Ser. A 356, 375–394 (1977). 57. E. Abramson, R. W. Field, D. Imre, K. K. Innes, J. L. Kinsey, J. Chem. Phys. 80, 2298–2300 (1983). 58. M. V. Berry, Proc. R. Soc. London Ser. A 413, 183–198 (1987). 59. E. P. Wigner, SIAM Rev. 9, 1–23 (1967). 60. V. Oganesyan, D. A. Huse, Phys. Rev. B Condens. Matter 91, 021001 (2019). 64. J. Bordé et al., Phys. Rev. Lett. 45, 14–17 (1980). 65. V. V. Albert, J. P. Covey, J. Preskill, Phys. Rev. X 10, 031050 (2020). 66. C. Zhang, P. G. Wolynes, M. Gruebele, Phys. Rev. A 105, 033322 (2022). 67. L. R. Liu et al., Supplementary data for ergodicity breaking in rapidly rotating C60 fullerenes, Harvard Dataverse (2023); https://doi.org/10.7910/DVN/E4KHWX [accessed 13 June 2023]. AC KNOWLED GME NTS We gratefully acknowledge comments on the manuscript from A. W. Young and Y.-C. Chan. Funding: This research was supported by AFOSR grant no. FA9550-19-1-0148; the National Science Foundation Quantum Leap Challenge Institutes (grant QLCI OMA- 2016244); the National Science Foundation (grant Phys-1734006); and the National Institute of Standards and Technology. Support is also acknowledged from the US Department of Energy, Office of Science, National Quantum Information Science Research Centers, Quantum Systems Accelerator. D.R. acknowledges support from the Israeli council for higher education quantum science fellowship and is an awardee of the Weizmann Institute of Science–Israel National Postdoctoral Award Program for Advancing Women in Science. P.J.D.C. and N.Y.Y. acknowledge support from the AFOSR MURI program (FA9550-21-1-0069). D.J.N. acknowledges support from the US DOE (DE-FG02-09ER16021) and NSF (CHE 2053117). T.V.T. acknowledges support from the NSF EPSCoR RII Track-4 Fellowship no. 1929190. Author contributions: L.R.L., D.R., and J.Y. designed, discussed, and performed the experiment and analyzed the data. L.R.L., D.R., P.B.C., P.J.D.C., D.J.N., N.Y.Y., T.V.T., and J.Y. participated in theory discussions and calculations and in writing of the paper. Competing interests: None declared. Data and materials availability: Data for raw absorption spectra and unwrapped tensor energy defect plots are deposited in Harvard Dataverse (67). All other data needed to evaluate the conclusions in the paper are present in the paper or the supplementary materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.sciencemag.org/about/ science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi6354 Supplementary Text Figs. S1 to S3 Table S1 Reference (68) Submitted 9 May 2023; accepted 23 June 2023 10.1126/science.adi6354 27. J. P. Aldridge et al., J. Mol. Spectrosc. 58, 165–168 (1975). Mater. Phys. 75, 155111 (2007). Liu et al., Science 381, 778–783 (2023) 18 August 2023 6 of 6
10.1126_science.adi1024
RES EARCH NEUROSCIENCE Climbing fiber multi-innervation of mouse Purkinje dendrites with arborization common to human Silas E. Busch1 and Christian Hansel1* Canonically, each Purkinje cell (PC) in the adult cerebellum receives only one climbing fiber (CF) from the inferior olive. Underlying current theories of cerebellar function is the notion that this highly conserved one-to-one relationship renders Purkinje dendrites into a single computational compartment. However, we discovered that multiple primary dendrites are a near-universal morphological feature in humans. Using tract tracing, immunolabeling, and in vitro electrophysiology, we found that in mice ~25% of mature multibranched cells receive more than one CF input. Two-photon calcium imaging in vivo revealed that separate dendrites can exhibit distinct response properties to sensory stimulation, indicating that some multibranched cells integrate functionally independent CF-receptive fields. These findings indicate that PCs are morphologically and functionally more diverse than previously thought. S1 and S2) and did not depend on zonal pattern- ing by zebrin II expression (15) (fig. S3). Physical constraints, however, might play a role in the spatial distribution of PC dendrite morphologies by foliar subregion (7). Indeed, the multiple dendrites of Split and Poly PCs predominantly ramified in a horizontal orientation (defined by <30° angle deviation from the PC layer, fig. S4, A and B; see materials and methods) in the sulcus of human folia (80%) while this occurred much less frequently in the bank and gyrus (23 and 25%, respectively; fig. S4D). This effect was not present in mice (fig. S4, C and D). In human PCs—where dendritic size expands strongly (Fig. 1, A and D)—the horizontal orientation follows the inward curvature of the sulcus, pos- sibly indicating a developmental response to physical constraints on growth. Because the physiological implications cannot be readily studied in humans, we turn to the corresponding mouse cells for further characterization. I nputs to the cerebellar cortex are integrated by the dendrites of Purkinje cells (PCs), its sole cortical output neuron. Despite their well-characterized position in what is consid- ered a conserved and stereotypical circuit (1), PCs exhibit diverse dendritic morphology in rodents (2) and it is not known how specific features of dendritic arborization may affect their function. Human PC morphology remains even more elusive. Studies of human PC morphology, which date back more than 120 years to the iconic illustrations of Golgi and Ramón y Cajal (3, 4), typically investigate small numbers of cells (5–7). Although no quantitative infor- mation on frequency and distribution of mor- phological types is available, it can be observed that human PCs are often “multibranched,” having either numerous trunks emerging from the soma or a proximal bifurcation of a single trunk. These features produce highly segregated dendritic compartments, raising the question of whether this confers functional properties that have gone unreported. We specifically asked whether the existence of several primary dendrites enables multiple climbing fiber (CF) innervation in the adult cerebellum. During development, the early growth of a primary dendrite provides struc- tural support for the ramification of a “winner” CF amidst competitive elimination of surplus CFs (8–10). Weaker CF inputs fail to translocate to the dendrite, possibly as a result of compet- itive processes resembling adult bidirectional synaptic plasticity (11–14). In PCs where mul- tiple primary dendrites conceivably offer a means to evade competition from other CFs, is the elimination pressure reduced enough to allow multiple CFs to be maintained? Would multi-innervation provide functionally inde- 1Department of Neurobiology and Neuroscience Institute, University of Chicago, Chicago, IL 60637, USA. *Corresponding author. Email: chansel@bsd.uchicago.edu pendent receptive fields to distinct dendri- tic compartments? A majority of human, but not murine, PCs have multiple primary dendrites Multiple CFs may innervate separate primary dendrites We used fluorescent calbindin immunolabel- ing to visualize PCs in postmortem human tissue (Fig. 1A). Based on proximal primary dendrite structure, which articulates the con- tours of the entire arbor, we define one stan- dard structural category—Normative, in which one primary dendrite may have a distant bifur- cation (beyond a two-somatic diameter thresh- old of 40 mm in mice and 50 to 70 mm in humans)– and two multibranched categories—Split, in which there is one trunk that bifurcates into multiple primary dendrites proximal to the soma (below the somatic diameter threshold) and Poly—, in which multiple trunks emerge directly from the soma (Fig. 1A and fig. S1; see mate- rials and methods). Although these categories translate to mice (Fig. 1D), we found that mice diverged significantly from humans in that they had fewer Split PCs (35.9 versus 44.8%) and far fewer Poly PCs (16.6 versus 51.2%; Fig. 1G and fig. S2A). Instead, in mice Normative PCs con- stituted the largest PC category (47.5%) in con- trast to humans (4.0%). We manually marked the distribution of dendritic morphologies of collectively ~8000 cells across whole parasagittal reconstructions of brain slices from the mid-hemisphere in hu- mans and mice (Fig. 1, B and E, and fig. S2, B and C). In posterior lobules of humans, there is a higher percentage of Poly PCs (53.8 versus 40.9%) and a lower percentage of Normative PCs (3.5 versus 6.1%) than in anterior lobules (Fig. 1C, fig. S2A, and table S1). Although the total rate is far lower, Poly PCs are relatively more prevalent in posterior lobules of mice as well (21.2 versus 10.3%; Fig. 1, F and G, fig. S2A, and table S2). The broad morphological distributions were consistent across nonpathological human and mouse individuals (fig. S2, B and C, and tables CF activity causes complex spike firing in PCs (16, 17), which is reciprocally related to simple spike firing (17, 18) and exerts powerful control over dendritic integration and PF plasticity (19–26). Though many studies cite the critical importance of one-to-one CF to PC connec- tivity in cerebellar function, as well as abnormal connectivity in dysfunction, some work has shown CF multi-innervation in ~15% of PCs in adult rodents (27–29). To test whether multiple CF innervation can be found in mature PCs, we combined a sparse dextran tracer (DA-594) labeling of inferior olivary (IO) neurons (Fig. 2A) with immuno- labeling of CF terminal boutons (VGluT2) and PCs (calbindin). Because all CF terminals are marked by VGluT2, but only some will express DA-594, this method allows for the identifica- tion of multiple CF inputs from distinct IO neu- rons onto single PCs (9, 30). Figure 2B shows a Poly PC (P87) that was indeed innervated by two CFs on its separate primary dendrites. Quantification of CF multi-innervation in mature PCs We obtained a quantitative measure of CF multi-innervation across the PC population by using whole-cell patch-clamp recordings in murine cerebellar slices (Fig. 2C). We adjusted current intensity and stimulus electrode posi- tion in the granule cell layer—subadjacent to each patched PC—to identify any ascending CF inputs and their stimulus thresholds. Mono- innervated PCs had a single, discrete excitatory post-synaptic current (EPSC) amplitude whereas multi-innervated PCs exhibited two or more dis- crete EPSC amplitudes selectively evoked by distinct stimulus intensities (Fig. 2C, bottom, and fig. S5, A and B). About 15% of all PCs in mature animals (P20-66) received multiple CFs Busch et al., Science 381, 420–427 (2023) 28 July 2023 1 of 8 A e v i t a m r o N t i l p S l y o P D RES EARCH | R E S E A R C H A R T I C L E B IV Intraculminate V Primary VI Superior Posterior VIIAfolium Crus I Horizontal III Preculminate Anterior Posterior 2 = 170.18 p < 0.001 C s s s l l l l l l e e e c c c f f f o o o t t t n n n e e e c c c r r r e e e P P P 80 80 80 60 60 60 40 40 40 20 20 20 0 0 0 L3 L4 L5 L6 CrusI L3 L4 L5 L6 CrusI L3 L4 L5 L6 CrusI CrusII CrusII CrusII L7B L8A L8B L7B L8A L8B L7B L8A L8B E VI Human Mouse 50µm 50µm N S P IV/V III II 2 = 106.83 p < 0.001 F s l l e c f o t n e c r e P 80 60 40 20 0 VII 500µm VIII IX X L2 L3 L4 L5 L6 CrusI CrusII L8A L8B L9 L10 VIIAtuber Crus II Ansoparamedian *** *** *** *** *** *** *** *** VIIB *** — — — — — — — — — — — — — — — — — — — — N one lobule S — — — — P Percents of total cell counts 47 36 17 45 51 5 mm G 80 80 80 Mouse Human 60 60 60 s s s l l l l l l e e e c c c f f f o o o t t t n n n e e e c c c r r r e e e P P P 40 40 40 20 20 20 0 0 0 2 = 55.27 p < 0.001 M H 4 Fig. 1. Comparative morphology and regional variability in human and mouse cerebellar PCs. (A) Immunolabeling of PCs in humans reveals a range of dendritic morphologies, categorized by primary dendrite geometry as Normative, Split, or Poly. (B) Human mid-hemisphere reconstruction demonstrating the spatial distributions of each morphological type. As a result of variable preservation of tissue some anterior lobules and intervening posterior sub- lobules had a lower density of labeled PCs. (C) Morphology demographics across lobules (n = 3 individuals >86 years old, 6640 cells; see table S1). (D) PCs filled with dye during a patch experiment in mice to scale with human cells also exhibit Normative, Split, and Poly morphology. (E and F) as in (B and C), but in mice (n = 3 mice >P50, 1350 cells; see table S2). (G) Morphological category distribution counted by lobule (top) in human (n = 20, 21, and 21 lobules) and mouse (n = 30, 30, 29) reveals a consistent increase in the number of Split and Poly PCs in human matching the rates of the whole cell population (bottom). Average lines depict median lobule value. *P < 0.05, **P < 0.01, ***P < 0.001. (Fig. 2D, left). CF competition for survival is complete by P20 (8, 9, 28, 31). In keeping with this, we did not find an effect of age on the rate of multi-innervation (fig. S6L). Combining this technique with fluorescent dye loading and confocal imaging revealed that multi-innervation was largely restricted to PCs with multi-branched structures (23/24 PCs) and occurred in ~25% of cells in this group (1/64 Normative, 15/61 Split, and 8/34 Poly PCs; Fig. 2D). The summed CF EPSC of multi-innervated PCs was larger, on average, than the amplitude of individual CF inputs to mono-innervated PCs (Fig. 2E). The amplitude of the smaller CF (at −30 to −10 mV holding potential) was typically >200 pA (Fig. 2E). This indicates that, under physiological membrane potentials, even the weakest of multiple CFs will likely deliver sufficient current to the soma to influence out- put (32). The amplitude of weaker CFs in- creased with age (fig. S5M), which may denote a delayed or elongated maturation period of these inputs relative to the completed develop- ment of single CF inputs or the more dominant of multiple CFs (fig. S5N). The relative EPSC am- plitude ratio between dominant and smaller CFs varied widely, but smaller CFs most often had Busch et al., Science 381, 420–427 (2023) 28 July 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E A B Cerebellum (Cb) Tracer (+) VGluT2 Tracer (-) VGluT2 Calbindin DA-594 VGluT2 25µ m Inferior Olive (IO) C 25µ m D 100 135 63 25µ m 2 = 15.71 p = 0.003 46 26 ) % ( s l l e C 75 50 25 0 22 2 1 2 3 14 1 1 1 2 3 1 2 3 Number of CFs 7 1 1 2 3 A p 0 0 5 20ms Mono-CF (1) Multi-CF (2+) Sum 2+ Weak * Mono-innervation 1 CF EPSC Multi-innervation 2 CF EPSCs A p 0 0 5 20ms F Split distance ( m) *** 00000 11111111111111111111111111111111111111111111111 1 11110111111111111111111111111111 111 111 10 1111111111111111111111111111111111111111111111100000 111111111111111111111111 111111111101111 100 G 1 CF 2+ CFs 77 7 7 7 75 50 0 0 55 5 5 0 5 5 5 22 2 2 25 0 0 0000000 0 0000000 0 0 ) m ( i e c n a t s d h c n a r B 150 100 50 0 E 0.0 ) A n ( C S P E F C -0.5 -1.0 -1.5 -2.0 ** * ** ) A p ( C S P E F C 0.0 -0.5 -1.0 -1.5 -2.0 H ) ° ( e l g n a k n u r t - y o P l 150 100 50 0 1 2+ 1 2+ Fig. 2. CF multi-innervation of mature multibranched PCs. (A) Schematic of tracer (DA-594) injection. (B) A Poly PC after immunolabeling for PCs (calbindin) and CF terminals (VGluT2). The tracer label distinguishes CFs with distinct olivary origin on the left and right trunks. (C) Scheme of whole-cell patch- clamp in cerebellar slices and CF EPSCs recorded from either a mono- or multi-innervated PC. (D) Number of mono- versus multi-innervated PCs as a combined population (left). Categorizing by morphology reveals that effectively all multi-innervation occurs in multibranched PCs (n = 50 animals, 159 cells). (E) Summed multi-CF EPSCs are larger than mono-CF EPSCs (n = 135 and 24 cells). The weaker of multiple CFs typically provides >200 pA signals. Holding potential: −10 to −30mV (n = 24 cells). (F) Multi-innervated PCs have earlier dendrite bifurcations (n = 135 and 24 cells). (G) Among PCs with a bifurcated primary dendrite, multi-innervated cells have a wider distance between compartments (n = 85 and 9 cells). (H) Multi-innervated Poly PCs have a wider angle between emerging trunks (n = 26 and 8 cells). Summary points indicate mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001. >25% the relative amplitude of the dominant CF (fig. S5O). This ratio differed across foliar sub-areas (fig. S5P) and correlated with the angle between Poly PC trunks (fig. S5Q), further em- phasizing the relationship between morphol- ogy and CF input properties. The prevalence of multi-innervation was correlated with proximity of bifurcation and angle of separation between emerging trunks in Split and Poly PCs, respectively (Fig. 2, F and H, and fig. S5, D and F). Multi-CF PCs also had wider dendritic arbors in the parasagittal plane (Fig. 2G and fig. S5G), but did not differ in the angle of bifurcation (fig. S2E) or soma size (fig. S5H). CF multi-innervation was present across cere- bellar regions and foliar sub-areas (fig. S6, A to I). Posterior lobules had a higher frequency of multi-innervation (fig. S6I), possibly due to increased prevalence of Poly PCs (fig. S6, H and J), matching our finding in immunolabeled tis- sue (Fig. 1F and fig. S2A). We did not observe a preferential rate of multi-innervation within the sulcus as a general pattern [but see (29) for more detailed analysis within vermis]. CF multi-innervation produces heterogeneous Ca2+ signals across dendrites in vivo Do multiple converging CFs provide function- ally distinct inputs to a single PC? How would this affect dendritic signaling in vivo? To answer these questions, we examined whether CF multi- innervation produces heterogeneous Ca2+ sig- nals across separate dendritic branches. CF input triggers massive Ca2+ entry into PC den- drites through voltage-gated Ca2+ channels (33), NMDA receptors (34), and release from internal stores (35), which can be locally modulated by ion conductance plasticity (33, 36) and inter- neuron inhibition (37, 38). These mechanisms contribute to the calcium events that we moni- tor here in vivo and to their modulation (39–41). We obtained a sparse PC expression of the Ca2+ indicator GCaMP6f and used two-photon imaging of mice in a state of quiet wakefulness (awake; no detected motion) to record non- evoked “spontaneous” Ca2+ signals from primary dendrite compartments in small populations of <10 cells (Fig. 3, A and B; see materials and methods). Volumetric scans visualized cellular morphology and permitted manual tracing of compartment ROIs (Fig. 3, B and D, and Figs. 4A and 5A) so fluorescence signals were ex- tracted and deconvolved separately to contrast event amplitude and frequency across branches (Fig. 3, C and E). In this configuration, non-evoked Ca2+ signals beyond the micro-compartment scale are almost entirely CF-dependent (42, 43) and Ca2+ event amplitude reflects the number of spikes in the presynaptic CF burst (44). This is confirmed by our observed ~1.2 Hz spontaneous Ca2+ event frequency (fig. S7D) that matches an expected CF input frequency moderately greater than 1 Hz (17). Busch et al., Science 381, 420–427 (2023) 28 July 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E A 0.5% AAV-L7-CRE + 20% AAV-FLEX-GCaMP6f B Branch 1 Branch 2 20µm Simplex 513nm 920nm Crus I Crus II D G 20µm ) % ( s t n e v e l a c o L 100 75 50 25 0 E 1.5 1 0.5 0 1 2 0.5 B - 0 1 B 0.5 -1 p = 0.087 *** *** H p m A 2 B 1.000 2 R j d A 0.100 0.010 0.001 20µm 3s 3s Normative 1.5 0.5 F/F0 1 0.5 0 1 Split 0.5 F/F0 0.5 ** *** * I p < 0.05 p > 0.05 0 0.5 -1 2 = 9.8 p = 0.007 J t u o h t i w s l l e C ) % i ( p h s n o i t a l e r h c n a r b 50 40 30 20 10 0 ) v e d t s ( l e a c s p m A 0.5 0.4 0.3 0.2 0.1 0.5 F/F0 0.5 F/F0 2s 2 C 1.5 1 2s 0.5 0 1.5 1 2 0.5 B - 0 1 B -0.5 -1 -1.5 Poly Branch 1 Dominates Branch 2 Dominates F *** ) 0 F F / ( e d u t i l p m a n a e M R2 = 0.099 p < 0.001 l ) v e d t s ( e a c s p m A 0.4 0.2 0.0 K 1.00 0.75 0.50 0.25 0.00 Global Local R2 = 0.078 p = 0.002 N S P B1 Amp N S P 20 40 60 80 100120 Split distance ( m) 50 100 150 200 250 Dendrite width ( m) Fig. 3. Two-photon imaging in vivo reveals Ca2+ signal heterogeneity across PC dendrites. (A) Schematic of experimental preparation. (B) Example imaging plane and three-dimensional reconstruction of a Poly PC. (C) Spontaneous signal and deconvolved events (circles) by branch with difference trace below demonstrates heterogeneous global event amplitude scale and branch-specific events. (D and E) Another recording from a Normative and Split PC highlights homogeneous versus heterogeneous signaling. (F) Local events are moderately smaller than global events (n = 15 animals, 95 cells). (G) Branch-specific local events as a percentage of total events in each cell by morphology (n = 15 animals; n = 32, 55, and 8 cells). (H) Linear regressions on branch cross-correlation quantifies branch similarity (left). Model fit R2 values (right) reveals that cells with low branch signal similarity are predominantly Split and Poly PCs (n = 32, 55, and 8 cells). Bordered points indicate nonsignificant covariance. (I) Cells lacking detectable relationship using regression on interbranch amplitudes are all Split or Poly PCs. (J and K) Interbranch amplitude variation by Split distance (n = 105 cells) and total parasagittal width of the dendrites (n = 109 cells). Summary points indicate mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001. We first identified local Ca2+ peaks detected in only one branch of each cell (Fig. 3, C, F, and G and movie S1), which were moderately smaller than globally expressed events on average (Fig. 3F). We also compared the inter-event cross- correlation of Ca2+ events across branches, for which the fit and significance of a linear re- gression describes the interbranch covariation (Fig. 3, H and I). Most PCs had homogenous Ca2+ signals with linear inter-event covariance relationships across branches (Adj R2 > 0.1) and low numbers of local events (Fig. 3, G and H). However, some PCs exhibited Ca2+ signal heterogeneity char- acterized by a linear regression of inter-event covariation with low Adj R2 < 0.1 that was not significant (0, 15, and 38% of Normative, Split, and Poly PCs, respectively; Fig. 3, H and I) or a higher ratio of local events (17.4, 36.6, and 51%; Fig. 3G). High variability of inter-event ampli- tude scale between branches, another measure of heterogeneity, correlated with the bifurcation distance and total parasagittal dendritic width (Fig. 3, J and K, and fig. S7C). This further links heterogeneity to underlying morphological con- tours defined by primary dendrite geometry. Confirming that local events are the product of additional CF input, PCs with high local event rates had higher mean (fig. S7, G, H, and P) and maximum total event rates (fig. S7, L and M), producing a larger dynamic range (fig. S7, N and O). Our observations link the occurrence of local CF events to underlying morphological contours defined by primary dendrite geometry, although other factors, such as inhibition by molecular layer interneurons (MLIs), are likely to contribute as well (37, 42). CFs convey distinct whisker receptive fields to separate primary dendrites To identify CF receptive fields (RFs) and their lo- calization on PC dendrites, we took advantage of the discrete organization of whiskers as a sen- sory input array (45, 46). We anaesthetized ani- mals to stimulate untrimmed individual whiskers at 2 Hz for 50 s periods while recording Ca2+ activity of PCs in medial Crus I (Fig. 4A and fig. S8A; see materials and methods). Most PCs had Busch et al., Science 381, 420–427 (2023) 28 July 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E A B1 B2 2Hz stimulation of C3 whisker B 0.3 0.2 0.1 0 0.2 0.1 2 B 0 - 1 B 0.1 5s B1 B2 5s 0.1 F/F0 0.1 F/F0 ns *** ** s e s n o p s e r l a c o l n a e M r e k s i h w r e p 10 5 0 C E G N S P Global Response Lateralized Response B1 B2 B1 B2 ns * * Unresponsive Global Lateralized ) m ( e c n a t s i d t i l p S 90 60 30 0 ns * * N S P 2 = 6.73 p = 0.035 Unresponsive Global Lateralized 19 46 10 10 38 28 4 10 4 N S P R2 = 0.046 p = 0.005 eD c n e r e f f i d r e k s i h w - r e t n I 15 10 5 0 F 100 s l l e c f o t n e c r e P 75 50 25 0 H 100 ) % ( s t n e v e l a c o L 75 50 25 0 1 0 2 Receptive field (whiskers) 4 3 Fig. 4. Branch-specific whisker receptive fields produced by CF multi-innervation of multibranched PCs. (A) Schematic of the imaging configuration and whisker stimulation under anesthesia. (B) Sample traces and deconvolved events by branch during 50 s whisker stimulation. Each whisker is tested twice; data from both periods are combined. Responsiveness of one branch and not the other drives an enhanced local event rate in B1 during the stimulus period. (C) Mean number of local branch events in response windows during stimulus periods of each tested whisker (n = 13 animals P95-120, n = 33, 112, and 24 cells). (D) Difference in local event number between whiskers eliciting maximum and minimum local responses (n = 33, 112, and 24 cells). (E) Schematic of global versus lateralized responses. (F) Percentage of PCs by dendritic response profile and morphological category. Fewer Normative PCs have lateralized responses than multibranched PCs (n = 169 cells). (G) Cells with lateralized responses have shorter Split distances (n = 75, 42, and 52 cells). (H) Cells with more spontaneous local events respond to a higher number of whiskers (n = 151 cells). Summary points indicate mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001. only global events identically represented across primary dendrites (figs. S8, A2, and A3). How- ever, some PCs had high numbers of local events in response windows during the stimulus pe- riod (Fig. 4, B and C, and fig. S8 A4) that varied in magnitude between distinct whisker stimuli (Fig. 4D), indicating RF selectivity. Anaesthetized activity is sparsened, so re- sponses were determined using the z-scored response probability during the experimentally bootstrapped high-frequency stimulus (fig. S8, B to D). Comparing the z-scored response probability of each dendrite, we observed a “lateralized” response in some PCs, in which local events of one branch constituted a whisker response not observed in the other branch (Fig. 4, E and F). Nearly all lateralized responses arose in Split and Poly PCs (48/52 cells, 92%; Fig. 4F) and in PCs with more spontaneous signal heterogeneity, which map to Split and Poly PCs as in previous experiments (fig. S8F). Furthermore, PCs with lateralized responses had more proximal dendrite bifurcations than PCs with only global responses (18.73 mm versus 28.24 mm; Fig. 4G). Notably, PCs with higher rates of branch-specific spontaneous events also ex- hibited responses to more whiskers, denoting an integration of more whiskers into their RFs (Fig. 4H). This supports the hypothesis that heterogeneous signals represent distinct, con- verging RFs such that heterogeneous PCs are sampling more upstream RFs carried by func- tionally independent CF inputs. CF-induced branch-specific representations of stimulus modality in awake mice Although anesthesia provided excellent con- trol and precision for single whisker stimula- tion, even subanesthetic ketamine alters network activity (47). To confirm that PC primary den- drites can differentially represent CF RFs in a more naturalistic state, we exposed awake ani- mals to uni- and multisensory stimuli (Fig. 5A). As a major hub for sensory integration during associative learning, PC dendrites are an impor- tant model for how converging input profiles are represented across dendrites. The amplitude and duration of CF-induced dendritic Ca2+ spikes depend on stimulus strength (43, 48), which is reflected in CF burst behavior (44) and also on synaptic connectivity and weight of the CF input itself (49). We stimulated awake animals with light (488 nm, ipsilateral), sound (12 kHz tone, bila- teral), and peri-oral air puff (10 psi, ipsilateral) stimuli either alone or in multimodal combi- nations while recording response properties in PC primary dendrites. Sensory-evoked events, more than spontaneous, typically produced a global dendritic signal with consistent intertrial amplitude ratio between branches (Fig. 5B1). However, we also observed complex sensory- evoked bursts of CF input with heterogeneous amplitudes between branches (Fig. 5, B2 and Busch et al., Science 381, 420–427 (2023) 28 July 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E A Light 488nm B2 C B1 50µm Sound 12kHz Puff 10psi 0.5 F/F0 2s Global Response ) % ( s t n e v e l a c o l x a M Uni * 100 Multi *** 75 50 25 0 N S/P N S/P D L l a c o L t n e c r e P 50 LP 100 P LPS LPS LS L 0 C S LP SP P C LS S SP E Branch Response BR (%) = B1Local - B2Local Branch 1 Dominates B1 B2 B1/2Global + B1Local + B2Local Dendrite Modality Response Profile Profile Mean/Range Putative CF RF space L LPSP S P LP LS SP Mean Range More B1 Local BR +100% More B2 Local 0% No difference BR -100% F ) % ( e g n a r R B 125 100 75 50 25 0 G ** ) % ( n a e m R B 50 25 0 -25 -50 -75 Bilateral low mean high range Unilateral high mean high range Global low mean low range or H Range - Mean * ns ) % ( y t i l a r e t a l i b R B 100 50 0 Branch 2 Dominates Branch 2 Alone B1 1.5 1 0.5 0 1 0 -1 B2 1.5 1 0.5 0 1 0 -1 B3 1.5 1 0.5 0 1 0 -1 B4 1.5 1 0.5 0 1 0 -1 Fig. 5. Branch-specific multisensory receptive fields. (A) Scheme of imaging and sensory stimulation of awake animals. (B) Sample traces showing combinations of inter-branch responses to different stimulus modalities. (C) The maximum number of local events observed for a stimulus of any category in Normative vs Split and Poly (S/P) PCs (here and below: n = 12 animals, n = 24 and 38 cells). (D) The percentage of responses having a local component, regardless of branch identity, across control (C), uni-, or multisensory trials. Lines connect values for each PC. (E) Calculation of DBR (top) between the stimulus types most favoring opposite branches. DBR values of each modality are calculated for each cell (bottom, schematic points) to map the DBR profile across stimuli and identify the mean and range. (F) The range is more pronounced in S/P cells (n = 24, 38). (G) No group difference in ΔBR mean (n = 24, 38). (H) Response profile bilaterality is the subtraction of DBR mean from the range. S/P PCs exhibit more bilaterality due to high ranges and low means (n = 24, 38). Summary points indicate mean ± SEM. *P < 0.05, **P < 0.01, ***P < 0.001. B3) and either branch-specific responses alone (Fig. 5B4) or combined with a global response (Fig. 5B2). Whereas PCs with multiple primary dendrites [Split and Poly (S/P)] had similar total response probabilities as Normative PCs (fig. S9C), a larger share of responses were branch-specific in S/P PCs across stimulus modalities (Fig. 5, C and D, and fig. S9, A and B). To assess the relationship between uni- and multimodal stimuli, we identified the maximum branch-specific responses to stimuli of each category (fig. S9D), obtained the difference between uni- and multisensory maxima, and found an enhanced rate of local responses in S/P but not Normative PCs (fig. S9F). This re- vealed that multisensory stimuli could enhance the differential representation of CF RFs across primary dendrites in putatively multi-innervated PCs while failing to influence mono-innervated Normative PCs. While the previous analyses were blind to branch identity, we next asked how much the differential representation of each stimulus could favor one branch over the other (fig. S9E). We generated a D Branch Response (DBR) index for each stimulus modality by calculating the difference in branch-specific, local responses as a fraction of total responses (Fig. 5E, top). Absolute DBR indicates the reliability of local responses on either branch whereas the sign of the DBR indicates which branch over-represented the modality. This allowed us to generate a pro- file of branch-specific representation across all stimulus modalities, which could be quantified by the DBR mean and range for each cell (Fig. 5E, bottom). In this way, PCs could be distinguished as having one of three classes of multisensory response profiles: global, with identical repre- sentation across branches in all cases; unilateral, with one branch exhibiting a larger RF repre- sentation than the other; and bilateral, with both branches capable of differentially repre- senting unique stimulus modalities. On average, S/P cells had a wider range, de- noting branch-specific (e.g., unilateral or bilat- eral) representations that were more distinct across modalities (Fig. 5F and fig. S9, G to I). Cells for which only one branch exhibits local responses—that is, unilateral—would have both a large DBR range but also a DBR mean that de- viates from zero to favor that branch. To better characterize whether some PCs had bilateral representation profiles, we calculated the bilat- erality of the RF profile by subtracting the DBR mean from the range. The local responses of S/P cells, more than Normative, produced RF profiles wherein a larger percentage of local signaling produced bilateral representations across sensory modalities (Fig. 5H and fig. S9, G, H, and K). Collectively, this shows that PCs with multiple primary dendrites can differen- tially represent RFs of distinct CF inputs across their separate dendrites in awake, mature mice (for a summary of heterogeneous signaling Busch et al., Science 381, 420–427 (2023) 28 July 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E observed in our two-photon recordings across the three dendrite morphologies in mouse PCs, see fig. S10). Discussion We found that noncanonical CF multi-innervation of PCs does occur in the mature murine cere- bellum and is dependent on primary dendrite morphology. Nearly all observed multi-innervation occurred in neurons with multiple primary dendrites. Based on a quantitative categoriza- tion of >6000 PCs from three human brains, we report that this type of PC dendritic struc- ture is predominant in the human cerebellum. By contrast, we detected that only a minority of murine PCs fall into the Split or Poly cate- gory. Within these morphological groups, about 25% of PCs were innervated by two or more CFs in the mouse. Our two-photon recordings suggest that most multi-innervated PCs have the capacity for branch-specific CF signaling and have distinct CF RFs. Our data do not allow us to conclude that the same results would be found in hu- man PCs if such recordings were possible. However, they do describe a new motif in PC dendritic compartmentalization: separate den- dritic subfields with their own assigned CF inputs may emerge when early branching forms a multibranched architecture. Variable CF burst frequency and modulation of CF input ampli- tude by MLIs may further contribute to com- partmentalization (37, 42). CFs provide instructive signals in cerebellar function and plasticity (17) by encoding signals related to error (19, 50), sensory omission (51), as well as reward or reward-prediction (52, 53). Our findings constitute a substantial shift from the currently held belief that one CF innervates each PC. Instead, our observations suggest that one CF innervates each primary dendrite. The consequences for dendritic integration and, ultimately, the activation of target cells in the cerebellar nuclei (54) are potentially multi- fold. Here, we discuss those that immediately result from geometric considerations. Multi- branched structure often increases dendritic width in the sagittal plane (Fig. 3K), in some cases even opening a cleft between compart- ments (see Fig. 1, A and D, Figs. 3B and 4E, fig. S1, B and C, and fig. S4B). This configuration inevitably leads to a wider physical gap between innervating PF bundles and thus to a potential functional separation of the contextual infor- mation they provide (55, 56). It is therefore conceivable that PCs drive spike output from a multitude of contextual input combinations that expand with increased dendrite size and complexity. Multiple CF innervation that, as we describe here, occurs at an elevated rate in multibranched PCs may serve several critical purposes. First, it may enhance PC function as a supervised associative learning perceptron that optimizes synaptic weights (57) by provid- ing RF-matched CF inputs—and thus relevant errors and instructive signals—to the different PF inputs that convey specific contextual infor- mation (fig. S11A). In this way, the perceptron orchestrates synaptic weight optimization based on compartmentalized, rather than all-dendritic, instructive signals. Thus, in receiving multiple CF inputs, some PCs are permitted to fully capi- talize on diverse context representations surveyed by their multibranched architecture. Second, multiple CF innervation may enable more com- plex PC computations, such as multiplexing and conveying input from a wider array of sensory modalities (fig. S11B). In this scenario, individual multibranched PCs may pair differ- ent contexts presented by their PF inputs with instructive information collected from a multi- modal environment. In both the human and mouse cerebellum, multibranched PCs are more prevalent in the posterior cerebellar hemisphere, a region linked to cognitive and affective roles (58–60). Whether or not the multibranched architecture enables such complex functions through the gained computational power that is postulated here needs to be investigated in future studies. On the other hand, excessive CF-PC strength resulting from disrupted synaptic pruning and ectopic innervation of distal PC dendrites is linked to pathological dysfunction in autism model mice (27, 49) mouse and human essen- tial tremor (61). The preferential targeting of distinct primary dendrites by multiple CFs (Fig. 2B) may bring computational advantages while avoiding disruptive enhancement of CF inputs to individual dendritic compartments. RE FERENCES AND NOTES 1. R. Apps, R. Hawkes, Nat. Rev. Neurosci. 10, 670–681 (2009). 2. H. Nedelescu, M. Abdelhack, A. T. Pritchard, J. Neurosci. Res. 96, 1476–1489 (2018). 3. C. Golgi, Sulla fina anatomia del cervelletto umano (Archivio Italiano per le Malattie Nervose, 1874), vol. 2. 4. S. Ramón y Cajal, Histologie du système nerveux de l’homme and des vertébrés (Paris : Maloine, 1909). 5. T. Kato, A. Hirano, J. F. Llena, Acta Neuropathol. 68, 145–148 (1985). 6. E. D. Louis et al., Brain 137, 3142–3148 (2014). 7. J. C. Eccles, M. Ito, J. Szentágothai, The Cerebellum as a Neuronal Machine (Springer, 1967). 8. K. Hashimoto, M. Kano, Neuron 38, 785–796 (2003). 9. K. Hashimoto, R. Ichikawa, K. Kitamura, M. Watanabe, M. Kano, Neuron 63, 106–118 (2009). 10. A. M. Wilson et al., Cell Rep. 29, 2849–2861.e6 (2019). 11. C. Hansel, D. J. Linden, Neuron 26, 473–482 (2000). 12. L. W. J. Bosman, H. Takechi, J. Hartmann, J. Eilers, A. Konnerth, J. Neurosci. 28, 798–807 (2008). 13. G. Ohtsuki, T. Hirano, Eur. J. Neurosci. 28, 2393–2400 (2008). 14. C. Piochon, M. Kano, C. Hansel, Nat. Neurosci. 19, 1299–1310 (2016). 15. N. L. Cerminara, E. J. Lang, R. V. Sillitoe, R. Apps, Nat. Rev. Neurosci. 16, 79–93 (2015). 16. J. C. Eccles, R. Llinás, K. Sasaki, J. Physiol. 182, 268–296 (1966). 17. J. I. Simpson, D. Wylie, C. I. De Zeeuw, Behav. Brain Sci. 19, 384–398 (1996). 18. A. Badura et al., Neuron 78, 700–713 (2013). 19. M. Ito, M. Sakurai, P. Tongroach, J. Physiol. 324, 113–134 (1982). 20. S. S.-H. Wang, W. Denk, M. Häusser, Nat. Neurosci. 3, 1266–1273 (2000). 21. V. Lev-Ram, S. T. Wong, D. R. Storm, R. Y. Tsien, Proc. Natl. Acad. Sci. U.S.A. 99, 8389–8393 (2002). 22. J. F. Medina, W. L. Nores, M. D. Mauk, Nature 416, 330–333 (2002). 23. M. Coesmans, J. T. Weber, C. I. De Zeeuw, C. Hansel, Neuron 44, 691–700 (2004). 24. C. Piochon et al., Proc. Natl. Acad. Sci. U.S.A. 113, 13221–13226 (2016). 25. M. L. Streng, L. S. Popa, T. J. Ebner, J. Neurosci. 37, 1997–2009 (2017). 26. H. K. Titley, M. Kislin, D. H. Simmons, S. S.-H. Wang, C. Hansel, J. Physiol. 597, 4387–4406 (2019). 27. C. Piochon et al., Nat. Commun. 5, 5586 (2014). 28. M. Kano et al., Neuron 18, 71–79 (1997). 29. H. Nishiyama, D. J. Linden, J. Neurosci. 24, 3926–3932 (2004). 30. T. Miyazaki, M. Watanabe, Anat. Sci. Int. 86, 10–18 (2011). 31. M. Kano, T. Watanabe, N. Uesaka, M. Watanabe, Cerebellum 17, 722–734 (2018). 32. I. Llano, A. Marty, C. M. Armstrong, A. Konnerth, J. Physiol. 434, 183–213 (1991). 33. M. T. Schmolesky, J. T. Weber, C. I. De Zeeuw, C. Hansel, Ann. N. Y. Acad. Sci. 978, 359–390 (2002). 34. C. Piochon, C. Levenes, G. Ohtsuki, C. Hansel, J. Neurosci. 30, 15330–15335 (2010). 35. H. Takechi, J. Eilers, A. Konnerth, Nature 396, 757–760 (1998). 36. G. Ohtsuki, C. Piochon, J. P. Adelman, C. Hansel, Neuron 75, 108–120 (2012). 37. J. C. Callaway, N. Lasser-Ross, W. N. Ross, J. Neurosci. 15, 2777–2787 (1995). 38. M. J. M. Rowan et al., Neuron 99, 999–1015.e6 (2018). 39. K. Kitamura, M. Häusser, J. Neurosci. 31, 10847–10858 (2011). 40. Y. Zang, S. Dieudonné, E. De Schutter, Cell Rep. 24, 1536–1549 (2018). 41. C. J. Roome, B. Kuhn, Nat. Commun. 9, 3388 (2018). 42. F. Najafi, A. Giovannucci, S. S.-H. Wang, J. F. Medina, Cell Rep. 6, 792–798 (2014). 43. M. A. Gaffield, A. Bonnan, J. M. Christie, Neuron 102, 762–769. e4 (2019). 44. S.-E. Roh et al., eLife 9, e61593 (2020). 45. L. W. J. Bosman et al., J. Physiol. 588, 3757–3783 (2010). 46. C. Ju et al., J. Physiol. 597, 2483–2514 (2019). 47. J. Cichon et al., Nat. Neurosci. 26, 39–52 (2022). 48. F. Najafi, A. Giovannucci, S. S.-H. Wang, J. F. Medina, eLife 3, e03663 (2014). 49. D. H. Simmons et al., Biol. Psychiatry Glob. Open Sci. 2, 450–459 (2021). 50. Y. Yang, S. G. Lisberger, Nature 510, 529–532 (2014). 51. S. Ohmae, J. F. Medina, Nat. Neurosci. 18, 1798–1803 (2015). 52. W. Heffley, C. Hull, eLife 8, e46764 (2019). 53. D. Kostadinov, M. Beau, M. Blanco-Pozo, M. Häusser, Nat. Neurosci. 22, 950–962 (2019). 54. A. L. Person, I. M. Raman, Nature 481, 502–505 (2011). 55. C. D. Wilms, M. Häusser, Nat. Commun. 6, 6464 (2015). 56. T. M. Nguyen et al., Nature 613, 543–549 (2023). 57. N. Brunel, V. Hakim, P. Isope, J.-P. Nadal, B. Barbour, Neuron 43, 745–757 (2004). 58. C. J. Stoodley, J. D. Schmahmann, Cortex 46, 831–844 (2010). 59. A. Badura et al., eLife 7, e36401 (2018). 60. A. LeBel, S. Jain, A. G. Huth, J. Neurosci. 41, 10341–10355 (2021). 61. M.-K. Pan et al., Sci. Transl. Med. 12, eaay1769 (2020). 62. S. Busch, Climbing fiber multi-innervation of mouse Purkinje dendrites with arborization common to human, Dryad (2023); https://doi.org/10.5061/dryad.kh18932c1. AC KNOWLED GME NTS For valuable advice and technical support, we thank Hansel lab members T. F. Lin, A. Silbaugh, and T. Pham. We thank R. A. Eatock and P. Mason (UChicago Neurobiology) for insightful discussions. For crucial feedback on the manuscript, we thank W. Wei and M. Sheffield (UChicago Neurobiology) as well as S. S. Wang (Princeton). M. Sheffield and S. S. Wang also provided preliminary viral tools. Human tissue made available by the Anatomical Gift Association of Illinois. Funding: This work was supported by the following: National Institutes of Health (NINDS) grant R21NS124217 (to C.H.); National Institutes of Health (NINDS) grant F31NS129256 Busch et al., Science 381, 420–427 (2023) 28 July 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E (to S.E.B.); and The University of Chicago Pritzker Fellowship (to S.E.B.) Author contributions: S.E.B. and C.H. designed the experiments. S.E.B. performed the investigation, formal analysis, visualization, and wrote the original draft. S.E.B. and C.H. acquired funding and edited the text. C.H. supervised the work. Competing interests: Authors declare that they have no competing interests. Data and materials availability: All data are available in the manuscript or the supplementary materials. Data and post-processing analysis code are publicly available in a Dryad archive (62). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. sciencemag.org/about/science-licenses-journal-article-reuse Figs. S1 to S11 Tables S1 to S10 References (63–66) MDAR Reproducibility Checklist Movie S1 SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi1024 Materials and Methods Submitted 3 April 2023; accepted 16 June 2023 10.1126/science.adi1024 Busch et al., Science 381, 420–427 (2023) 28 July 2023 8 of 8
10.1126_science.adi8237
RES EARCH R E S E A R C H A R T I C L E S U M M A R Y ◥ NEUROSCIENCE Associative and predictive hippocampal codes support memory-guided behaviors Can Liu†, Ralitsa Todorova†, Wenbo Tang, Azahara Oliva, Antonio Fernandez-Ruiz* INTRODUCTION: The brain generates models of the environment that are used to guide flexible behaviors. This process requires learning the states of the world (such as specific locations) as well as the transitional relationships be- tween those states (e.g., successive locations along often-traveled trajectories). The hippo- campal cognitive map is believed to be one such internal model, supporting a variety of behaviors, including associative learning, nav- igational planning, and inference. It remains unknown which facets of hippocampal coding are required for these different behaviors and how they support both associative and predic- tive memory functions. RATIONALE: We hypothesize that two modes of hippocampal activity support learning of world states and state transitions, respectively. On one hand, the synchronous coactivity of groups of hippocampal neurons—cell assemblies—may encode features of individual states, forming an associative code. On the other hand, the ordered activation of these cell assemblies into behaviorally relevant sequences may encode the relational structure between states, form- ing a predictive code. Previous research has not been able to dissociate these two dynamic codes or provide evidence of their specific functions. We leveraged an optogenetic ap- proach to dissociate these two coding schemes, with the goal of disrupting the predictive code (hippocampal sequences) while preserving the associative code (rate coding and coactivity dynamics) in behaving rats. This dissociation allowed us to examine the different memory functions of these two codes. RESULTS: We optogenetically perturbed the fine temporal coordination of hippocampal place cell firing as rats navigated specific spa- tial trajectories in a novel maze. This manip- ulation disrupted properties of the predictive code (such as temporally compressed place cell sequences and anticipatory place field shifts), but global network dynamics and single-cell spatial tuning and rate coding properties were preserved. During sleep after the novel experience, we observed that task-related cell assemblies en- Associative code Predictive code Optogenetic perturbation Associative code Predictive code Coactivity, theta sequences Reactivation, replay Coactivity, No theta sequences Reactivation, No replay Unaffected by perturbation n.s. e c n a m r o f r e P Disrupted by perturbation *** e c n a m r o f r e P Flexible navigation Associative learning Associative and predictive codes in the hippocampus. Our optogenetic manipulation perturbed hippocampal sequences without affecting cell coactivity, thus selectively disrupting the predictive code. After learning, cell assemblies were reactivated, but their order was not preserved, abolishing sequential replay (right). Perturbing the predictive code had no effect on associative learning (bottom left) but did disrupt the flexible learning of novel optimal trajectories on a maze (bottom right). [Rat illustrations: Yu Kang] coding discrete maze locations were reacti- vated in sharp wave–ripples (SWRs), un- affected by the manipulation. However, their sequential structure did not reproduce the order in which they were active in the task, resulting in impaired sequential replay for the perturbed trajectories. This result shows a dissociation between assembly reactivation and sequence replay, two phenomena previously assumed to reflect the same underlying pro- cess. The same manipulation did not disrupt replay of familiar trajectories, suggesting that the precise temporal coordination of place cell firing during learning mediates initial plasticity required for subsequent replay. Computational simulations suggest that distinct Hebbian plas- ticity mechanisms mediate assembly reactiva- tion and sequence replay. We tested the functional role of the predic- tive code by deploying our optogenetic manip- ulation in two different hippocampal-dependent memory tasks. Context-reward associative learn- ing in a conditioned place preference task was unaffected and thus does not require a predic- tive map or memory replay. On the other hand, flexible memory–guided navigation in a forag- ing task was perturbed by the manipulation and thus depends on hippocampal predic- tive coding. CONCLUSION: Our results provide a mechanis- tic and functional dissociation between coac- tivity and sequence codes in the hippocampus. Hippocampal cells with similar responses to behavioral variables fire together, forming func- tional assemblies during learning, which are reactivated in SWRs during subsequent sleep. These cell assemblies encode discrete states in the environment, an associative code that is sufficient for some types of episodic memories. As these cell assemblies are activated in a spe- cific order during behavior, they form tempo- rally compressed hippocampal sequences and promote Hebbian plasticity. This process en- ables the replay of behaviorally relevant se- quences during SWRs. Hippocampal sequences thus encode transitional structures of world states, generating a predictive model on top of the associative code of individual assem- blies. This new framework contributes to our understanding of how memory associations develop into predictive representations of the world and helps reconcile previously dis- parate views on hippocampal function.▪ Department of Neurobiology and Behavior, Cornell University, Ithaca, NY, USA. *Corresponding author. Email: afr77@cornell.edu †These authors contributed equally to this work. Cite this article as C. Liu et al., Science 382, eadi8237 (2023). DOI: 10.1126/science.adi8237 READ THE FULL ARTICLE AT https://doi.org/10.1126/science.adi8237 Liu et al., Science 382, 283 (2023) 20 October 2023 1 of 1 RES EARCH R E S E A R C H A R T I C L E ◥ NEUROSCIENCE Associative and predictive hippocampal codes support memory-guided behaviors Can Liu†, Ralitsa Todorova†, Wenbo Tang, Azahara Oliva, Antonio Fernandez-Ruiz* Episodic memory involves learning and recalling associations between items and their spatiotemporal context. Those memories can be further used to generate internal models of the world that enable predictions to be made. The mechanisms that support these associative and predictive aspects of memory are not yet understood. In this study, we used an optogenetic manipulation to perturb the sequential structure, but not global network dynamics, of place cells as rats traversed specific spatial trajectories. This perturbation abolished replay of those trajectories and the development of predictive representations, leading to impaired learning of new optimal trajectories during memory-guided navigation. However, place cell assembly reactivation and reward-context associative learning were unaffected. Our results show a mechanistic dissociation between two complementary hippocampal codes: an associative code (through coactivity) and a predictive code (through sequences). L earning associations from experience and using this knowledge to make novel pre- dictions are integral features of memory- guided behavior. Learning the structure of one’s environment can be conceptual- ized as two separate processes: learning the states in the world and learning the likely transitions between those states. The hippo- campus has been implicated in both processes because it has been shown to be fundamental for encoding spatial-temporal associations (1–5) and supporting flexible memory–guided behaviors, planning, and inference (6–11). Such involvement in diverse types of memory has led to different theories of hippocampal function focused either on its role in the formation and recall of episodic memories (12–15) or in the generation of predictive models to guide behav- ior (10, 16, 17), while some proposals high- lighted the intrinsic relationship between these functions (18, 19). However, it is unclear which facets of hippocampal coding are required for these different behaviors. One prominent candidate is the short-timescale (~10 to 30 ms) coincidental firing of hippocampal neurons with selective responses to external variables (e.g., place cells encoding the same location). This synchronous activation of “cell assemblies” can lead to the strengthening of their connec- tions through Hebbian plasticity, forming a coactivity code (20–22). In addition, hippocampal cell assemblies organize into sequences. During navigation, temporally compressed neuronal sequences on the timescale of a single cycle of the theta oscillations (6 to 12 Hz; i.e., “theta sequences”) sweep ahead of the animal’s cur- rent position to encode future routes (23–25). During rest periods, the sequential activation of place cells “replays” past spatial trajectories (26, 27). Coactivity and sequential hippocam- pal dynamics have been traditionally consi- dered conjoined aspects of the same under- lying process, supported by common circuit and synaptic mechanisms to the extent that they are regarded as nondissociable. An alternative possibility is that coactivity and sequential dy- namics of hippocampal place cells represent distinct coding schemes (Fig. 1). First, assem- blies of coactive cells would encode individual associations, or discrete states in the world. Second, assemblies can be concatenated into sequences reflecting either experienced or in- ferred behaviorally relevant state transitions, forming a predictive code. Although multiple studies have found deficits in memory-guided behavior by disrupting hippocampal activity (28–32), none have been able to dissociate these two dynamic modes and provide evidence of the specific role of hippocampal sequences in behavior. To overcome this limitation and test our hypothesis, we selectively disrupted hippo- campal sequences in rats while preserving place cell expression and coactivity dynamics and investigated their specific contribution to predictive coding and flexible spatial learning. Results Disruption of hippocampal theta sequences impairs the formation of a predictive map To dissociate assembly coactivation and the sequential organization of assembly activity, we sought to specifically disrupt place cell sequence dynamics while preserving their tuning and coactivity properties in the main hippocampal output region, CA1. Timing of CA1 place cell firing relies on inputs from the me- dial entorhinal cortex (MEC) (33–35). However, silencing or inactivating the MEC profoundly Department of Neurobiology and Behavior, Cornell University, Ithaca, NY, USA. *Corresponding author. Email: afr77@cornell.edu †These authors contributed equally to this work. A B C D Fig. 1. Two hypothesized memory codes in the hippocampus. (A) As a rat learns to navigate a maze to find a reward, place cells activate along its spatial trajectory (color areas represent place fields). (B) During navigation, but also during offline periods, place cells with overlapping fields are active together, forming functional assemblies (top), and these assemblies form a sequence that reproduces an animal’s learned trajectory (bottom). (C) We propose that the coactivity of cell assemblies represents an associative code to learn discrete states in the world, while their sequential activation forms a predictive code to learn behaviorally relevant state transitions. (D) An associative memory code may be sufficient for some types of memory, such as associating a place with a reward. Other types of memory-guided behavior, such as flexibly navigating a complex environment, would require a predictive model. Liu et al., Science 382, eadi8237 (2023) 20 October 2023 1 of 16 RES EARCH | R E S E A R C H A R T I C L E interferes with hippocampal physiology, affect- ing place cell expression, coactivity, and phase coding, and thus lacks the necessary specificity to test the role of sequences (36–39). To over- come these limitations, we expressed channel- rhodopsin (ChR2) in rat MEC g-aminobutyric acid–releasing (GABAergic) cells to optogen- etically entrain them at an artificial gamma frequency (40, 41). With this approach, we selectively perturbed the fine temporal struc- ture of MEC outputs while simultaneously recording hippocampal population activity (Fig. 2A and fig. S1). In contrast to previous manipulations (33, 36), our optogenetic pertur- bation spared theta power, frequency, and the theta phase modulation of CA1 unit firing (fig. S1). To evaluate the effect of our optogenetic perturbation on CA1 place cell properties, we delivered light while rats (n = 5) ran in one direction (“Stim ON”) but not in the opposite direction (“Stim OFF”) on a novel linear track. We aimed to disrupt specific sequences that encoded part of the experience and to leave other sequences intact as a control. The forma- tion, stability, and firing properties of CA1 place cells at the single cell and population levels were preserved for both the Stim ON and Stim OFF directions (Fig. 2, B to F, and figs. S2 to S4). Place cells in the two directions had similar properties and formed a highly direc- tional code such that Stim OFF and Stim ON trajectories could be decoded separately (fig. S3A), and on the population level, the compo- sition, spatial properties, and stability of place cell assemblies were similar in the Stim OFF and Stim ON directions (figs. S4, E to H). Despite preserved place cell tuning curves, the progressive shift in their spike timing in rela- tion to the phase of global theta rhythm [“phase precession” (42)] was impaired in the Stim ON direction (Fig. 2D and fig. S5). More- over, the sequential firing of pairs of place cells at the theta timescale was disrupted in Stim ON (Fig. 2, G and H). In contrast, across theta cycles, cofiring of CA1 cells was preserved in both directions (Fig. 2, G and I). In our framework, the formation of place fields (PFs) in a novel environment reflects learning the states in the world (e.g., loca- tions), whereas the formation of a predictive map develops as an animal learns the likely transitions between those states (e.g., succes- sive locations along often-traveled trajectories). To test whether the predictive code could be dissociated from the encoding of the individ- ual states with our manipulation, we examined the evolution of place cells and hippocampal predictive coding features in the novel maze. We found that place cell properties in both Stim OFF and Stim ON directions showed experience-dependent development. Place field quality in both directions improved across laps, as measured by both their spatial infor- mation (Fig. 3B) and the decoding accuracy of animal position from place cell population activity (fig. S3), indicating that spatial maps were formed and developed with experience, unaffected by our optogenetic perturbation. We then examined the development of a predictive code in two ways. We first analyzed theta sequences, because they are temporally compressed representations of spatial trajec- tories, reflecting short-term predictions of the immediate future (43–46). In contrast to the preserved spatial map formation, theta se- quences were impaired in the Stim ON direc- tion (Fig. 3A). Moreover, we observed a specific experience-dependent development of theta sequences in the Stim OFF direction, such that they became more prospective across laps (en- coding longer distances ahead of the animal) (Fig. 3C), a process that could reflect the for- mation of increasingly predictive internal rep- resentations of the maze. This development was not observed in the Stim ON direction (Fig. 3C). Another property of the hippocampal predictive map is the progressive backward shift of place fields toward past states (or loca- tions) with experience (47), so that the early firing reflects the expectation of upcoming states (17). We observed a progressive backward shift across laps of place fields only in the Stim OFF direction, as expected, not the Stim ON direc- tion (Fig. 3, D and E). These changes in the pre- dictive code were not due to differences in animal behavior between the two trajectories (fig. S3G). Impaired replay with preserved reactivation of a novel experience Memory formation in the hippocampus is typ- ically measured as both the reactivation of short-timescale coactivity patterns (“assem- blies”) and the replay of neuronal sequences during sharp wave–ripples (SWRs) (48–50). We first examined sequential replay of a running experience on a novel linear track. During pauses between track running and during sleep after the task, we observed SWR-associated replay events reproducing place cell sequences of the Stim OFF direction (Fig. 4, A and B). However, replay sequences of the Stim ON trajectory, where hippocampal predictive coding was dis- rupted, were impaired (Fig. 4B). Compared with baseline sleep sessions before the task, the fraction of SWRs that expressed signifi- cant replay of the Stim OFF trajectory and their sequence scores were considerably increased, but no experience-dependent enhancement was observed for the Stim ON trajectory (Fig. 4C and fig. S6). This effect cannot be explained by differences in single neuronal activity during SWRs or decoding quality (fig. S7). Stim ON replay sequences were not observed in either the forward or the reverse direction (fig. S6, B and C), and this result was replicated with different replay methods and decoding param- eters (figs. S6 and S7). We also tested whether this disruption of replay of specific trajectories was due to the impaired formation of a predic- tive map (i.e., encoding the relevant state tran- sitions). We conducted an experiment in which rats (n = 3) first explored a linear track without manipulation to form a predictive map with- out perturbation and, in a subsequent session, experienced optogenetic perturbation in one direction as previously described. Theta se- quences were still impaired in the Stim ON direction, but we observed intact replay for both Stim ON and Stim OFF trajectories (fig. S8), confirming that although predictive cod- ing is key in the initial stages of encoding novel experiences, once the predictive map is established, it does not require further repeti- tions to generate replay of recent experience. In the novel linear track experience, we next examined reactivation of task coactivity pat- terns by detecting synchronous cell assemblies that encoded discrete positions along both Stim OFF and Stim ON trajectories (fig. S9). The composition, spatial properties, and sta- bility of assemblies were similar in the Stim OFF and Stim ON directions (fig. S4, E to H). Surprisingly, Stim OFF and Stim ON assem- blies reactivated with similar strength during SWRs (Fig. 4, A and B), significantly increasing their activity compared with the preceding sleep (Fig. 4D and fig. S9, B to D). This result indi- cates an intact memory reactivation of loca- tions along both trajectories, which was also confirmed using other canonical metrics of memory reactivation (Fig. 4E and fig. S9A). We wondered how these stark differences in reactivation and replay can coexist. A closer look at the temporal dynamics of cell assem- blies revealed an important difference. Assem- blies from the Stim OFF trajectory reactivated during SWRs in a temporal sequence that re- produced the order of their encoded spatial locations in the maze (Fig. 4 and fig. S9A). However, while individual assemblies for the Stim ON trajectory reactivated during SWRs with similar strength to those for the Stim OFF trajectory, their task sequential structure was not preserved in these reactivation events (Fig. 4B and fig. S9E), resulting in the impair- ment of sequence replay and the preservation of reactivation of the Stim ON trajectories. In- deed, although the strength of assembly reac- tivation and replay for the Stim OFF direction were correlated at the level of individual SWRs, this correlation was absent for the Stim ON direction (Fig. 4F), dissociating reactivation from replay in the absence of a predictive code. Associative and predictive codes support different memory-guided behaviors Next, we asked whether the associative (coac- tivity) and predictive (sequences) hippocam- pal codes have different functional roles. A predictive model of the environment would be required to learn tasks that necessitate flexible navigation. To test this prediction, we trained Liu et al., Science 382, eadi8237 (2023) 20 October 2023 2 of 16 RES EARCH | R E S E A R C H A R T I C L E A E F 800 600 400 200 ) F F O m i t S y b d e r e d r o ( D I l l e c B C G Stim OFF Stim ON D Stim OFF 800 600 400 200 ) N O m i t S y b d e r e d r o ( D I l l e c 1 m min max firing rate Stim ON H density 14× mean 0 100 I ) s m ( g a l e m i t 50 0 -50 -100 100 50 0 -50 -100 ) s m ( g a l e m i t -100 0 field distance (cm) 001 -100 0 field distance (cm) 001 Fig. 2. Optogenetic perturbation of MEC gamma timing impairs temporal but not rate coding features of CA1 place cells. (A) Histological sections showing (top) expression of mDlx-ChR2-mCherry in MEC GABAergic cells (red) and optic fibers’ tracks (dashed lines) and (bottom) probe locations in CA1. (B) Rate maps for CA1 pyramidal cells (n = 815) in the Stim OFF and Stim ON directions. (C) Example Bayesian decoding of CA1 population spike trains during running behavior. Dashed cyan line represents animal’s actual position. The color map indicates probability of a decoded location. (D) Representative place cells for the Stim OFF (top) and Stim ON (bottom) directions. (Top) Firing rate as a function of spatial positions. (Bottom) Theta phase–position raster of place cell spikes. Red line represents the linear circular regression. (E) Spatial information was not significantly different between cells in the Stim OFF and Stim ON directions (P = 0.39; Wilcoxon signed-rank test). (F) Representation stability as measured by the population vector (PV) correlation between even and odd trials in both directions was higher than shuffled data (OFF/ON versus shuffle: P = 6 × 10−8/P = 6 × 10−8, Wilcoxon signed-rank test) and not different between the ON and the OFF directions (P = 0.35, Wilcoxon signed-rank test). (G) Examples of theta firing relationships for cell pairs with overlapping place fields in OFF (top) and ON (bottom) directions. The cells cofired in the same theta cycles in both cases, but the order of firing (AB) is only consistent in the OFF direction. (H) Average normalized density plot between place field distance and theta timescale firing lag for all overlapping place cell pairs in OFF (top) and ON (bottom) running directions. Theta compression was disrupted in Stim ON only (P = 9 × 10−5, Wilcoxon signed-rank test between theta compression slopes; see methods). Circles indicate the cell pairs shown in (G). (I) Cofiring of overlapping place cells in the same theta cycles was preserved in both directions (overlapping versus nonoverlapping place cells in OFF/ON: P = 1 × 10−4/P = 9 × 10−5, Wilcoxon signed-rank test; OFF versus ON: P = 0.067). ***P < 0.001. rats (n = 5) to learn a novel spatial configura- tion of rewards in a familiar “cheeseboard” maze (8, 41, 51). Rats learned to find three hidden water rewards in different locations on the maze, with the reward locations changing every day (Fig. 5, A and B). Each trial started with the animal in the home box and ended when they retrieved the three rewards and came back to the box. The most efficient strat- egy for the animal in this task was thus to find the optimal route that connected all three re- ward locations by learning to sequentially Liu et al., Science 382, eadi8237 (2023) 20 October 2023 3 of 16 RES EARCH | R E S E A R C H A R T I C L E A E B C D Fig. 3. Impaired development of place cell theta sequences and predictive properties. (A) (Top) Example theta sequences. (Bottom) Average (n = 20 sessions in five rats) decoded position estimated across different theta phases, relative to the actual animal position (dashed black line). Stim ON theta sequences were impaired (Stim OFF versus Stim ON quadrant scores: P = 7 × 10−141; weighted correlation: P = 4 × 10−22, Wilcoxon rank sum test, n = 17,992/13,265 theta cycles). (B) Increase in spatial information of place cells across laps in both Stim ON (blue) and Stim OFF (black) conditions (***P = 1.8 × 10−11/1.2 × 10−5 for Stim OFF/ON, Wilcoxon rank sum tests comparing the 1st versus the 15th lap). (C) Increase in theta sequence look-ahead index in the Stim OFF (***P = 4.0 × 10−6), but not the Stim ON, direction (P = 0.71, Wilcoxon signed-rank tests). (D) Backward shifting of PF center-of-mass (COM) overlaps on the Stim OFF trajectory (***P = 1.3 × 10−69) but not the Stim ON trajectory (P = 0.16, Wilcoxon signed-rank tests). COM shift relative to the first lap is shown (>0, forward shifting to future; < 0, backward shifting to past). (E) Example place cells showing backward shifting place fields on the Stim OFF trajectory (top) and stable fields on the Stim ON trajectory (bottom). Arrows depict the animal’s running direction. Curves on top show smoothed firing rate for the first and last laps, and raster plots below show spikes across all laps. For all these analyses, n = 20 sessions from five rats were used. navigate from one reward to the next. On half of the days, we performed the same optoge- netic perturbation during learning trials (Stim ON sessions), whereas in the other half there was no stimulation (Stim OFF sessions). We observed an intact formation of place fields (fig. S10) and impaired theta sequences (Fig. 5D) in Stim ON sessions. Animal learning per- formance was impaired in Stim ON compared with Stim OFF sessions (Fig. 5C and fig. S10), supporting a role of the predictive code in the efficient learning of novel optimal trajecto- ries in a familiar environment. We next asked whether replay would be preserved despite the disruption of theta sequences in this familiar environment. We first examined replay of the novel trajectory learned by the end of the cur- rent day’s training session. In Stim OFF sessions, we found robust replay in both awake SWRs (during intertrial periods in the home box) and in the post-learning sleep session (Fig. 5F and fig. S11A). In contrast, replay of the novel route to the new goal locations was impaired in Stim ON sessions (Fig. 5F and fig. S11A), suggesting that when learning a new optimal route to goals, even in familiar environments, the tran- sitional structure between locations of the route encoded by theta sequences is required for its subsequent replay. Replay of the familiar routes to previously learned locations was pre- served in both Stim OFF and Stim ON sessions (fig. S11B). These effects were not due to dif- ferences in SWR properties, single-unit firing dynamics during SWRs, or decoding accuracy of animal position between Stim OFF and Stim ON sessions (figs. S11, D and E, and S10C). The two-dimensional (2D) place field properties of individual cells were also preserved, similar to those on the 1D linear track (fig. S10E). In contrast to the impaired Stim ON replay, reactivation of cell assemblies was maintained in both Stim OFF and Stim ON sessions (Fig. 5G and fig. S11), replicating our results from the linear track. Finally, we tested whether im- pairment of hippocampal sequences with pre- served place cell expression and coactivity affected subsequent memory recall. Memory recall measured 2 and 22 hours after learning was disrupted in Stim ON compared with Stim OFF sessions (Fig. 5H), and animals spent less time exploring around the reward locations after Stim ON than after Stim OFF sessions (fig. S10B). Furthermore, while performance on the 2- and 22-hour tests during Stim OFF ses- sions remained on par with the final perform- ance reached during learning, in Stim ON sessions, performance was further deteriorated (fig. S10A). Our hypothesis predicted that in contrast to “model-based” behaviors that require a pre- dictive code, an associative memory code im- plemented by the coactivity of hippocampal cells would be sufficient to support other types of episodic memory, such as associating spatial context with rewards. We trained rats (n = 5) on a conditioned place preference (CPP) paradigm (Fig. 5I), which depends on intact hippocampal function. Rats were trained to associate one of the two halves of a cheeseboard maze with the presence of water rewards at random locations. After four pairing trials, we tested their pref- erence for either of the two halves of the maze (without reward present). In half of the ses- sions, we delivered the optogenetic perturba- tion during all running periods during pairing trials. In line with our previous results, we de- tected similar assembly reactivation during sleep after both Stim OFF and Stim ON pairing Liu et al., Science 382, eadi8237 (2023) 20 October 2023 4 of 16 RES EARCH | R E S E A R C H A R T I C L E A B C D F E 0 -0.5 0 0.5 5 0 5 0 0 1 0 1 Fig. 4. Impaired replay with preserved reactivation of a novel experience. (A and B) Examples of replay and reactivation for (A) the Stim OFF and (B) the Stim ON experience, respectively. (Top) Decoded position during replay events (color coded by decoded position). (Middle) Raster plots of neuronal firing; colored circles indicate activations of assemblies composed of neurons with nearby place fields (colored ticks; colors of assemblies reflect their peak activation position). (Bottom) Assembly reactivation strength curves. (C) Increase in proportion significant replay (left) and sequence scores (right) of SWRs in post-task sleep as compared with pre-task sleep. Unlike the OFF direction, there was no significant replay in the ON direction (n = 4437 events; Stim OFF: student’s t test, ***P = 7 × 10−11; Stim ON: P = 0.29 for proportions; and Wilcoxon signed-rank test: OFF ***P = 3 × 10−17 and ON P = 0.29 for scores). Only the first session in the maze was included (n = 5 sessions). (D) Reactivation strength of task-related assemblies centered on post-task sleep SWR, increased relative to baseline (Stim OFF post-task sleep versus baseline sleep during SWRs: Wilcoxon signed-rank test, n = 31 components, **P = 0.0096; Stim ON post-task sleep versus baseline sleep during SWRs: Wilcoxon signed-rank test, n = 27 components, **P = 0.0026). (E) Reactivation of pairwise neuronal correlation as measured by explained variance (EV) in post-task sleep was significant for both directions (Stim OFF: *P = 0.031; Stim ON: Wilcoxon signed-rank test, *P = 0.031; Stim OFF versus Stim ON, Wilcoxon signed-rank test, P = 0.81; n = 5 sessions). (F) Correlation between replay scores and assembly reactivation strength for post-task SWRs computed independently for Stim OFF and Stim ON replay events. Reactivation and replay were correlated in Stim OFF (Pearson’s correlation coefficient r = 0.11, ***P = 3 × 10−12) but not in Stim ON (Pearson’s r = −0.028, P = 0.06). trials (fig. S12). Population activity during SWRs formed distinct memory representations for both contexts (rewarded and nonrewarded maze halves) in both Stim OFF and Stim ON sessions, consistent with an intact associative code (fig. S12). Because the CPP task involves boundaries, it gave us the opportunity to quan- tify a critical signature of hippocampal predic- tive coding, which is the elongation of place fields along obstacles in 2D mazes, reflecting the topological structure of the space (17, 52). Similar to previous reports (53, 54), fields near the boundary of our maze tended to elongate parallel to the wall in Stim OFF sessions (Fig. 5K and fig. S12). However, this effect was dis- rupted during Stim ON (Fig. 5K and fig. S12), providing further support that the predictive map was impaired. To assess replay in the ab- sence of stereotypical trajectories, we exam- ined the degree to which the sequential order of firing in post-task sleep SWRs resembled the sequential order in behavior. The sequen- tial order was preserved in post-task sleep in Stim OFF sessions but not in Stim ON sessions (fig. S12, F and G). Despite these changes in the predictive code, on the test session, where no reward or stimulation was delivered, rats showed an enhanced preference for the pre- viously rewarded context in both Stim OFF and Stim ON conditions (Fig. 5, J and L), con- firming that an associative memory code was sufficient for the formation and retrieval of a context-reward association. Circuit mechanisms for coactivity and sequence memory codes Given this dissociation, we wondered which distinct circuit mechanisms support the reac- tivation and replay of memories. The main inputs that modulate the firing dynamics of CA1 pyramidal cells are excitatory projections from CA3 and the entorhinal cortex as well as perisomatic inhibition (Fig. 6A). The rela- tive strength of each of these inputs during behavior can be estimated by measuring layer- specific gamma oscillations (34, 35, 55, 56). MEC optogenetic perturbation specifically disrupted midfrequency gamma oscillations (gammaM) in the distal CA1 dendrites, the domain in- nervated by MEC axons (Fig. 6B and fig. S13), while other CA1 gamma oscillations, elicited respectively by CA3 inputs and local inhibition (34, 35, 55–57), remained unaltered (Fig. 6B and fig. S13). The coupling of CA1 spikes to gammaM, but not to the other two gamma oscillations, was also impaired (fig. S13), con- sistent with the notion that our manipulation disrupted CA1 pyramidal cell entrainment by MEC inputs but not by CA3 or local interneur- ons. The disruption of theta sequences during initial experience led to the impairment of re- play but not the reactivation of assemblies, raising the question of which mechanisms link these processes. To answer this question, we turned to computational modeling. Previous work has suggested that the highly recurrent CA3 hippocampal region with its auto-associative network structure is respon- sible for both the coactivation of CA1 cells forming functional assemblies (22, 58) and the generation of sequential activity (24, 59, 60). We implemented a spiking CA3-CA1 network model capable of generating spontaneous re- play. Our CA3 network consisted of recurrent- ly connected leaky integrate-and-fire pyramidal Liu et al., Science 382, eadi8237 (2023) 20 October 2023 5 of 16 RES EARCH | R E S E A R C H A R T I C L E A B I D E C H F G Stim ON/Stim OFF baseline 15 min sleep pairing 4 x 20 min sleep testing 15 min J Stim OFF Stim ON K L n i e m T i ) % ( e n o z d e d r a w e r 100 *** *** 50 0 baseline test Stim OFF Stim ON test baseline Fig. 5. Internally generated sequences are needed for memory-guided behavior. (A) Task structure of the cheeseboard task: Each day, animals learned a novel trajectory to get three hidden rewards in a circular arena. (B) Schematic of the cheeseboard setup. (C) Learning performance measured as distance traveled relative to optimal trajectory. ANOVA with repeated measures showed a significant main effect of group (F1,43 = 133.4, P < 10−10; n = 33 and 11 sessions for Stim OFF and Stim ON, respectively). (D) Average decoded position versus theta phase, relative to actual animal position (dashed black line). Stim ON theta sequences were degraded (quadrant scores were lower in Stim ON than in Stim OFF: n = 5958/2695 cycles for Stim OFF and Stim ON, respectively, P = 2 × 10−3; weighted correlations were lower in Stim ON than in Stim OFF: P = 3 × 10−4, rank sum test). Two cycles are shown for visibility. (E) Example of a decoded replay event. Decoded linearized position (top), spike raster (middle), and assembly reactivation strength (bottom) were shown for the same replay event. (F) Increase in replay score in post-task sleep as compared with pre-task sleep. Unlike the Stim OFF condition, there was no significant replay in the Stim ON condition (Stim OFF: P = 3 × 10−20; signed-rank test, n = 11,460 events; Stim ON: P = 0.41, n = 8689 events). (G) Reactivation strength of task-related assemblies centered on post- task sleep SWR (Stim OFF post-task sleep versus baseline sleep during SWRs: n = 36 components, P = 4.71 × 10−4 ; Stim ON post-task sleep versus baseline sleep during SWRs: n = 22 components, P = 3.19 × 10−3, signed-rank test). (H) Memory performance during 2 and 22 hour post-learning recall tests (P = 6.8 × 10−4/7.9 × 10−4 for 2 and 22 hour tests, rank sum test; n = 25/11 for Stim OFF/ON sessions from five rats). (I) Task structure of CPP task. (J) Example rat paths for Stim OFF (left) and Stim ON (right) baseline and testing sessions on CPP task. The side rewarded during pairing is highlighted in yellow. (K) Example place fields near the boundary for Stim OFF (left) and Stim ON (right) conditions, illustrating place field elongation along the boundary, which was disrupted in the Stim ON condition. Red circles emphasize the shape of the place field. (L) CPP memory performance: both Stim ON and Stim OFF training resulted in animals spending more active time in rewarded versus unrewarded side (Stim OFF: paired t test, n = 5 sessions, P = 2.85 × 10−4; Stim ON: paired t test, n = 6 sessions, P = 2.10 × 10−5). cells and interneurons (61). This network was connected in a feedforward manner to a CA1 network, with a similar composition but lack- ing recurrent excitation, so every CA1 pyram- idal cell integrated inputs from multiple CA3 pyramidal cells (Fig. 6C). We first simulated a learning phase, akin to a rat running on a novel linear track, in which place cells received random spatial inputs. In accordance with experimental observations, CA3-CA1 synaptic weights were updated according to an asym- metric (62, 63) spike time–dependent synap- tic plasticity (STDP) rule, whereas CA3-CA3 weights followed a symmetric (64) STDP rule (Fig. 6C). On the basis of our experimental findings, we simulated a Stim OFF condition, in which place cells displayed phase preces- sion and robust theta sequences (Fig. 6D and fig. S14). In a Stim ON condition (mimicking the effects of our optogenetic MEC perturba- tion), place cells were phase-locked to theta oscillation but lacked phase precession and theta sequences (Fig. 6D and fig. S14), repro- ducing our experimental findings. We also simulated offline epochs [akin to non–rapid eye movement (non-REM) sleep] before and after spatial learning; in such periods, the network received low-level stochastic spiking inputs to drive activity. We examined CA1 spiking patterns after learning and quanti- fied replay and reactivation as we did with our experimental data. In the Stim OFF con- dition, synapses between overlapping place cells were potentiated by STDP (fig. S14). Place cell sequences experienced in the learning phase were spontaneously replayed in the fol- lowing offline epoch (Fig. 6, E and F). In the Stim ON condition, disorganized timing within theta cycles disrupted CA3-CA1 STDP, which relies on a consistent timing between pre- and postsynaptic spike pairs (fig. S14). CA3-CA3 STDP was preserved due to place cell coactivity within theta cycles, regardless of their precise temporal ordering (fig. S14). These effects re- sulted in an impairment of replay in the Stim ON condition (Fig. 6, E and F) but robust reactivation of place cell assemblies in both conditions (Fig. 6, E and G), mirroring our ex- perimental results. The dissociation between reactivation and replay in our model suggests that reactivation and replay are differently mediated by STDP at CA3 and CA1 synapses, providing a plausible circuit-level mechanism for our findings. Discussion In this study, we described two complemen- tary hippocampal circuit mechanisms that sup- port the formation of memory associations and the generation of predictive representations of the environment. First, the coactivity of place cell assemblies encoded discrete states in the envi- ronment and, with their subsequent reactivation, Liu et al., Science 382, eadi8237 (2023) 20 October 2023 6 of 16 RES EARCH | R E S E A R C H A R T I C L E A B C D F G E n o i t i s o p d e d o c e D r e t s a R g n i r i F n o i t a v i t c a e R h t g n e r t s ) s t i n u z ( 4 0 n o i t i s o p m 3 0 min max certainty 100 ms 4 0 -0.5 0 0.5 Fig. 6. Circuit mechanisms for coactivity and sequence hippocampal dynamics. (A) Schema depicting the inputs to CA1 pyramidal neurons stratified along their somatodendritic axis. Local inhibitory inputs are dominant in the pyramidal layer (st. pyr.), CA3 inputs target proximal apical dendrites in the stratum radiatum (st. rad.), and entorhinal inputs target distal dendrites in the stratum lacunosum-moleculare (st. l-m.). (B) Gamma amplitude-theta phase comodulograms for each layer-specific LFP component (see methods) displayed modulation in a specific gamma sub-band (averaged data from n = 10 sessions from four rats): CA1pyr in gammaF (100 to 160 Hz), rad in gammaS (30 to 60 Hz), and LM in gammaM (60 to 110 Hz). (Right) MEC perturbation selectively impaired LM gammaM but not CA1pyr or rad oscillations (**P = 0.002, Wilcoxon signed-rank test; n = 13 sessions from five rats). (C) (Left) Model schematic depicting a subnetwork of CA3 cells projecting to a subnetwork of CA1 cells. Triangles represent pyramidal neurons, and circles represent inhibitory interneurons. (Right) STDP rules used to train different synapses within the network during learning trials. (D) In Stim OFF simulations (top), place cells displayed phase precession and prominent theta sequences. In Stim ON simulations (bottom), phase precession was disrupted, and theta sequences were abolished. (E) Example replay events simulated by the model after Stim OFF (left) and Stim ON (right) learning. Decoded position, spike raster, and assembly reactivation strength as in Fig. 4. (F) Increase in proportion of “SWR” events with significant replay in post-task “sleep” epochs in Stim OFF and Stim ON protocols. Replay improvement was above chance levels in Stim OFF but not Stim ON simulations (Stim OFF: P = 0.0039, Wilcoxon signed-rank test, n = 6585 events; Stim ON: P = 0.19, n = 6031 events). (G) Reactivation strength of task-related assemblies centered on post-task sleep SWR-like events increase relative to baseline (Stim OFF post-task sleep versus baseline sleep during SWR-like events: Wilcoxon signed-rank test, n = 362 components, ***P = 5 × 10−44; Stim ON post-task sleep versus baseline sleep during SWR-like events: Wilcoxon signed- rank test, n = 312 components, ***P = 1 × 10−27). supported associative memory. Second, tempo- rally compressed hippocampal sequences en- coded transitional structures of states in the environment, supporting the formation of a predictive map. Replay of these sequences provided a mechanism to update and exploit this predictive map, but only after intact en- coding of transitional structures of task states during learning. Disruption of trajectory-specific theta sequences impaired their subsequent replay but did not affect the reactivation of neuronal assemblies representing those same locations. Reactivation and sequence replay are therefore dissociable neural processes. The dependence of replay on learning environ- ment transitional structures through theta se- quences was highly specific for both space and time: Neither disruption of theta sequences for the same trajectory traveled in the opposite direction nor disruption of previously expe- rienced trajectories affected replay. Further- more, these results indicate that encoding the specific transitions between place cells along behaviorally relevant trajectories, rather than place field formation per se or global theta- timescale dynamics alone, is an instrumental mechanism for replay, explaining and extending previous observations on the relationship between theta population dynamics and replay (65–67). The disruption of temporally compressed hippo- campal sequences impaired flexible memory– guided navigation but did not affect the for- mation and recall of contextual associations. Our results advance a framework to unify previously disparate views of hippocampal function, including encoding cognitive maps for spatial and episodic memory (68–70), pre- dictive maps for flexible navigation and plan- ning (10, 16, 17), and multimodal memory representations for relational processing (18, 71). We argue that the hippocampal role in encoding episodic memories relies on two complementary Liu et al., Science 382, eadi8237 (2023) 20 October 2023 7 of 16 RES EARCH | R E S E A R C H A R T I C L E mechanisms. First, the synchronous coactivity of cell assemblies binding different features of experience into cohesive representations enables the fast encoding of discrete states in the world and supports associative memory. Second, the formation of a predictive map is supported by internally generated hippocampal sequences, binding discrete successor states into a relational structure. Moreover, our results can account for two independent lines of evidence supporting previous theories of hippocampal predictive coding: predictive sequences and “successor representation” place fields. Our manipulation impaired the development of theta and replay sequences. Previous studies have shown that both theta and replay se- quences not only reproduce animal recent ex- periences but can also construct novel predictive representations. Theta sequences encode trajec- tories ahead of the animal’s current location and can represent alternative paths at decision points (44–46). Replay sequences can repro- duce never-experienced paths (72, 73) and dynamically change to reflect learned task contingencies (11, 74–76) or represent available paths not yet taken (77, 78). In both cases, internally generated hippocampal sequences could be conceptualized as a sampling process between a series of possible future states from a probabilistic generative model, and they were both impaired by our manipulation. In addition, experience-dependent backward ex- pansion and deformation around obstacles of place fields are believed to reflect the encoding of a predictive map over the locations the animal expects to occupy in the near future (17, 52). These features were disrupted by our manipulation, in contrast to place field forma- tion and stability, which may be supported by different plasticity mechanisms (79). Together, these effects of our manipulation suggest that all these “generative” features of hippocampal representations are linked by common circuit mechanisms. Although numerous processes could con- tribute to the generation of replay sequences (80–83), our results suggest that initial theta sequence–mediated STDP is necessary for the replay of newly learned spatial trajectories but not for memory reactivation. These results suggest a hierarchical organization of hippo- campal assembly and sequence dynamics. Previous work investigating the postnatal development of hippocampal dynamics sup- ports this dissociation. Synchronous cell as- semblies are already present around postnatal day 17 in rats, with their reactivation in SWRs reflecting encoding of individual locations on a maze (84). However, theta sequences and replay encoding animal spatial trajectories ap- pear only after postnatal day 21 (66, 84). The present findings agree with previous studies that investigated the contribution of the temporal coordinatization of hippocampal spike timing to memory. Pharmacological dis- ruption of place cell theta rhythmicity and co- ordination, while preserving their rate coding properties, impaired spatial memory (85, 86). In addition, disruption of the behavioral time- scale (on the order of seconds) sequential struc- ture of CA1 firing responses during temporal delays also impaired memory (31, 87). These studies highlight the importance of the precise temporal coordination of hippocampal spike timing for memory. Additional research is still needed to eluci- date the precise contribution of different input pathways and cell types in the entorhinal- hippocampal network to temporal coding and sequence generation. MEC inputs reach CA1 via direct projections from layer III and also indirectly via layer II inputs to the dentate gyrus, CA3, and CA2 (88). Our manipulation was not restricted to either of these input path- ways and therefore cannot precisely determine their contributions to the generation of CA1 sequences. Furthermore, a small subset of MEC GABAergic cells send direct long-range pro- jections to CA1 (89). It is possible that our manipulation also affected these projections; however, their sparsity and the fact that the stronger GABAergic projections to CA1 arise from the lateral rather than the medial en- torhinal cortex (90) limit their potential con- tribution to the results described here. Overall, this study suggests the coexistence of complementary associative and predictive codes in the hippocampus. This dual code could account for the wide range of behavioral func- tions attributed to this brain structure in learn- ing, memory, navigation, and planning. Materials and methods Surgical procedures Rats (adult male Long-Evans, 300 to 500 g, 3 to 6 months old) were kept in the vivarium on a 12 hour light/dark cycle and were housed two per cage before surgery and individually after it. All experiments conformed to guide- lines established by the National Institutes of Health and have been approved by the Cornell University Institutional Animal Care and Use Committee. Silicon probe implantation was performed as described previously (41, 51, 91). Animals were anesthetized with isoflurane anesthesia and craniotomies were carried out under ste- reotaxic guidance. Silicon probes (NeuroNexus, Cambridge Neurotech, or Diagnostic Biosig- nals) were mounted on custom-made 3D-printed microdrives to allow precise adjustment of the vertical position of sites after implantation. The probes were inserted above the target re- gion. Craniotomies were sealed with sterile wax. Two stainless steel screws were placed bilaterally over the cerebellum to serve as ground and reference electrodes. Several addi- tional screws were driven into the skull and covered with dental cement to strengthen the implant. Finally, a copper mesh mounted on a 3D-printed resin base was attached to the skull with dental cement and connected to the ground screws to act as a Faraday cage, attenuating the contamination of the recordings by envi- ronmental electric noise and protecting the headgear. Three doses of analgesic were administered, with the first dose administered prior to surgery in order to cover 72 hours total. After post-surgery recovery, probes were moved gradually in 50- to 150-mm steps per day until the desired position was reached. Hippo- campal layers were identified physiologically by unit activity and characteristic local field potential (LFP) patterns (34, 92). A variety of different silicon probes were im- planted in the dorsal hippocampus [−4.0 to 4.5 mm anteroposterior (AP) from Bregma and 2.6 mm from midline (ML)]. Data from some of the animals included in this study have also been included in a previous study, and surgical procedures have been described in more detail there (41). For optogenetic experiments, rats were also implanted with custom-made optic fiber arrays (three 200-mm core multimode fibers each, ~500 mm apart, connected to a single 2.5-mm steel ferrule; Doric Lenses) in both hemispheres over the MEC (–7.7/–8.4/–9.1 AP; ± 4.6 ML and 4.7/ 4.3/3.2 mm from the surface of the brain, for each fiber respectively). Optogenetic experiments For optogenetic experiments, rats were in- jected with custom-prepared AAV5-mDlx- hChR2(H134R)-mCherry from AddGene [plasmids were a gift from Dr. Gord Fishell (40)]. Three injections per hemisphere were performed along the dorsoventral MEC as follows: (i) −7.7 AP, ±4.6 ML, 4.7 mm depth, 200 nl; (ii) −8.4 AP, ±4.6 ML, 4.3 mm depth, 400 nl; (iii) −9.1 AP, ±4.6 ML, 3.2 depth, 700 nl. After injection, craniotomies were sealed, and animals recov- ered in the vivarium for 3 weeks. After this period, a second surgical procedure for im- planting optic fibers and electrodes was per- formed, as described above. Optic fiber arrays were implanted in the same craniotomies per- formed previously for virus injection. For opto- genetic stimulation, fiber array ferrules were connected with mating sleeves to 450-nm blue light-emitting laser diodes coupled to 2.5mm steel ferrules (PL-450, Osram). Optogenetic perturbations were performed by delivering blue light, modulated with a pos- itive 53 Hz current sinusoid using an isolated current driver (Thorlabs). Light intensity was calibrated for each animal during home cage recordings by analyzing the suppression of LFP gamma power in the stratum lacunosum- moleculare during stimulation. A minimum and maximum power of 3 and 6 mW, respec- tively, was used. In the linear maze task, light Liu et al., Science 382, eadi8237 (2023) 20 October 2023 8 of 16 RES EARCH | R E S E A R C H A R T I C L E stimulation was triggered when the animal crossed an infrared sensor near the end of the track (but before the reward port located at the very end of it) and stopped when it crossed a similar sensor at the other end, resulting in stimulation only during running periods. In the cheeseboard maze and CPP tasks, the same opto- genetic stimulation was applied during all runn- ing periods in the maze during learning trials. Behavioral and electrophysiological recordings After surgery, animals were handled daily and accommodated to the experimenter, recording room, and cables for 1 week before the start of the experiments. Before the start of the behav- ioral experiment, the animals were water- restricted. Electrophysiological recordings were conducted using Intan RHD2000 inter- face board or Recording Controller (IntanTech) and 64-channel digital headstages (IntanTech or Diagnostic Biochips), sampled at 20 kHz. For all behavior tasks, animal position was re- corded with an overhead camera (Basler) and tracked with DeepLabCut (93) and custom codes. Linear track task In the linear track, rats were trained to run back and forth to collect small sugar water re- wards at both ends. Animals typically per- formed between 30 and 80 trials per day. The session was terminated when the animal was satiated, typically after 40 min. The linear track was placed 1 meter above the floor and was 300 cm long and 7 cm wide with 5- to 10-cm-high walls. Water rewards were automatically de- livered in reward wells at both ends of the track. Baseline and post-task sleep were recorded in the home cage before and after the task, respectively. Cheeseboard maze task The cheeseboard maze was a circular platform (120 cm diameter), where the animals learned to find three goal wells that contained water rewards. A trial was completed once the ani- mal had retrieved all rewards and returned to the start box to collect an additional food pellet reward. The locations of the goal wells changed daily but were fixed during a given session. This strategy required the animals to update their memory for the new goal loca- tions in an otherwise familiar environment during each session. Note that between trials, there was always a delay of ∼30 s. If the rat could not find all three rewards within 2 min, the trial was terminated, and the animal was re- turned to the start box. A pre-task test session of five trials (with the same configuration as the previous day) was run every day to assess whether the animal remembered the previous day’s positions, followed by a 2-hour sleep ses- sion. Next, a 30-trial learning session was con- ducted (with or without light stimulation). After a 2-hour sleep session after the learning trials, a five-trial post-task test session (with the newly learned reward configuration) was also conducted to examine whether the animal remembered the newly learned locations. Learn- ing performance was evaluated as the distance traveled from the start box to collect the three rewards and divided by the optimal trajectory (shortest possible path). A value of 1 then indi- cates an optimal path taken by the animal. Performance in the post-task test sessions was computed in the same way (Fig. 5H), and ad- ditionally it was compared directly with the learning performance on the final learning trial (fig. S10A). An additional measure of perform- ance was computed as the proportion of ex- ploration time spent within 5 cm the correct wells (fig. S10B). Stim OFF and Stim ON ses- sions were alternated in a counterbalanced manner across animals. The inclusion of data for the analysis required that the animal be pretrained for a week in this task. Conditioned place preference task The CPP task was performed in the same maze as in the cheeseboard maze task but separated into two compartments by a 30-cm-high wall in the midline. Animals had free access to both compartments connected by a common corri- dor at one end of the maze (Fig. 5I). The task included three stages spanning 3 days: base- line, pairing, and testing. For the baseline stage on day 1, animals were allowed to freely ex- plore the two compartments for 15 min with no reward. A baseline place preference was then measured as the difference in active ex- ploratory time spent in the two compartments. The compartment that animals explored less was selected as the rewarded side for the sub- sequent pairing sessions. Pairing sessions in- cluded one session after the baseline session on day 1, two sessions on day 2, and one ses- sion before the testing session (see below) on day 3. During the pairing sessions, animals explored the maze for 20 min, with hidden water reward delivered at random locations on the rewarded side. On day 3, after the final pairing session, a testing session was per- formed, during which animals again explored the maze without reward for 15 min and the place preference was measured as was done during the baseline session. Optogenetic stim- ulation was delivered when animals actively explored the maze during pairing sessions. Each CPP session was flanked by sleep record- ings (around 2 hours) in the home cage. Learn- ing performance of the CPP task was measured by comparing the place preference during the baseline versus the testing session (Fig. 5L). Tissue processing and immunohistochemistry After the termination of the experiments, ani- mals were deeply anesthetized and perfused transcardially, first with 0.9% saline solution followed by 4% paraformaldehyde solution. The brains were sectioned into 70-mm-thick slices (Leica Vibratome). The sections were washed and mounted on glass slides with a fluorescence medium [Fluoroshield with DAPI (4′,6-diamidino-2-phenylindole), product no. F6057, Sigma, USA]. A confocal microscope (Zeiss LSM 800) was used to obtain high-quality photos. Computational model Architecture We built a simplified network model consist- ing of 1250 pyramidal cells (PYR) and 100 in- hibitory interneurons (INT) in area CA3 and 1250 pyramidal cells and 100 inhibitory inter- neurons in area CA1 using the Brian2 simu- lation environment (94). Each neuron was modeled as an adaptive exponential leaky integrate-and-fire unit with cellular adapta- tion (61). Briefly, Cm dV tð Þ dt ¼ (cid:2)ðgL V tð Þ (cid:2) Vrest ð Þ (cid:2) (cid:1) gLDT exp (cid:3) V tð Þ (cid:2) q DT þ Isyn tð Þ þ w tð ÞÞ where V(t) is the membrane potential, Cm is the membrane capacitance, gL is the leak con- ductance, Vrest is the reversal potential of the linear leak current, q is the spike threshold, DT is the threshold sharpness, Isyn is the synaptic current, and w(t) is the adaptation current, described by dw tð Þ dt ¼ a V tð Þ (cid:2) Vrest ð tw Þ (cid:2) w tð Þ where the parameter a describes the strength of the subthreshold adaptation. The synaptic current Isyn was computed as Isyn tð Þ ¼ gAMPA tð Þ V tð Þ (cid:2) Eexc ð Þ þ gGABA tð Þ V tð Þ (cid:2) Einh ð Þ where Eexc = 0 mV and Einh = −70 mV are the reversal potentials of excitatory and inhibitory currents, respectively. Synapses were modeled as conductances with biexponential kinetics (61). Synapses Within each area, PYR→INT connectivity oc- curred at 10% probability, and INT→INT and INT→ PYR connectivity occurred at 25% prob- ability. Whereas PYR→PYR connectivity was absent from CA1, CA3 PYR cells projected at 10% probability to PYR from both CA3 and CA1, mod- eling recurrent CA3 connectivity and CA3 in- puts into CA1, respectively. Synaptic parameters were used as previously reported (61), with the modification that synaptic weights were initialized to follow a lognormal distribution exp(N(0,1)) × winit, where wCA3→CA3 ¼ 0:3 and where wCA3→CA1 ¼ 0:7. To simulate learning as a result of the activity during exploration, STDP init init Liu et al., Science 382, eadi8237 (2023) 20 October 2023 9 of 16 RES EARCH | R E S E A R C H A R T I C L E was modeled for PYR→PYR synapses alone, where synaptic weights are updated according to an additive pair-based learning rule as follows (cid:3) (cid:1) Dwþ ¼ Aþ exp (cid:2) at tpost if tpre < tpost Dt tþ (cid:1) Dw(cid:2) ¼ A(cid:2) exp (cid:2) (cid:3) Dt t(cid:2) at tpre if tpre > tpost CA3→CA3 STDP followed a symmetric rule, where A+ = A− = 80 pA, and t+ = t− = 62.5 ms, while CA3→CA1 STDP followed an asymmetric rule, where A+ = 800 pA, A− = −A+ × 0.4, tþ ¼ 20 ms, and t(cid:2) ¼ 40 ms. All weights were cropped at wmax = 40 nS. Exploration phase Spike trains of PYR cells during exploration were generated as previously described (61). Briefly, 10% of all pyramidal cells in each re- gion were designated as place cells, and place fields were randomly distributed along a 3-m- long simulated linear track. Exploration was simulated as 10 min of exploration time while the rat ran along the linear track at 50 cm/s. Silent cells (nonplace cells) fired at 0.01 Hz, and place cell spikes were sampled from a Poisson process at 20 Hz, with the sampling procedure ensuring an inhomogeneous Pois- son process with a time-dependent rate l(t), which was the product of a Gaussian tuning curve (representing the neuron’s place field) with width s = 7 cm (yielding l = 30 cm place fields) and a theta component. The theta com- ponent was either a phase precession compo- nent to simulate the sequences of the Stim OFF condition or a phase-locking component to simulate the lack of sequences but preserved theta-cycle coactivity in the Stim ON condition. The Stim OFF theta component was (cid:1) qOFF tð Þ ¼ cos 2pfqt þ (cid:3) (cid:5) (cid:4) x tð Þ (cid:2) xi p 2p (cid:3) Q where Q represents the proportion of the linear track that the place field spans and takes numeric value of 30-cm field over 300-cm linear track or 0.1, and xi is the position of the place field of cell i. The Stim ON theta component was a phase- locking component designed to yield the same phase-locking value as the other condition’s theta precessing component but abolishing sequential order of spikes (cid:1) qON tð Þ ¼ cos 2pfqt þ (cid:3) (cid:7) p 2p (cid:3) Q (cid:6) N p 2 ; s l (cid:5) where l is the length of the track (3 m), and (cid:4) N p is a normal distribution with mean p 2 2 and standard deviation s l. ; s Offline phase Before and after exploration-triggered STDP, the network was simulated using Brian2 to respectively model pre (baseline) and post “sleep” sessions. To drive the network, each CA3 pyramidal cell received random inputs in the form of uncorrelated Poisson spike trains with a pooled mean rate of 15 Hz. Under these conditions, the network spon- taneously generated population burst events (akin to SWRs). Analysis All the modeling steps described above were performed n =10 times with different random- ly generated initial conditions determining place field locations, initial weights, etc. This resulted in n = 10 simulated sessions, in which theta sequences, replay, and reactivation were analyzed applying the same methodology and parameters as we did to analyze experimental data described above. Spike sorting and unit classification Spike sorting was performed semiautomatical- ly with KiloSort (95) (https://github.com/cortex- lab/; KiloSort), followed by manual curation using the software Phy (https://github.com/ cortex-lab/phy) and custom designed plug-ins (https://github.com/petersenpeter/phy-plugins) to obtain well-isolated single units. Cluster quality was assessed by manual inspection of waveforms and autocorrelograms, and by the isolation distance metric. Multiunit, noise clus- ters, or poorly isolated units were discarded for analysis. Well-isolated units were classified into putative cell types using the Matlab pack- age CellExplorer (96) (https://github.com/ petersenpeter/CellExplorer). Spiking char- acteristics, including the autocorrelograms, spike waveforms, and putative monosynap- tic connections derived from short-term cross- correlograms, were used to select and charac- terize well-isolated units. Three cell types were assigned: putative pyramidal cells, narrow waveform interneurons, and wide waveform interneurons. Two key metrics used for this separation were burst index and trough-to- peak latency. Burst index was determined by calculating the average number of spikes in the 3- to 5-ms bins of the spike autocorrelo- gram divided by the average number of spikes in the 200- to 300-ms bins. To calculate the trough-to-peak latency, the average waveforms were taken from the recording site with the maximum amplitude for the averaged wave- forms of a given unit. Detection of brain states State scoring was performed as previously de- scribed (41, 51). Briefly, the LFP was extracted from wide-band data by lowpass filtering (sync filter with a 450 Hz cutoff band) and down- sampling to 1250 Hz. Broadband LFP, nar- row-band theta frequency LFP, and estimated electromyogram (EMG) were used for state scoring. Spectrograms were computed from broadband LFP with a fast Fourier transform in 10-s sliding windows (at 1 s), and a principal components analysis was computed after a Z-transform. The first principal component reflected power in the low (<20 Hz) frequency range, with oppositely weighted power at higher (>32 Hz) frequencies. Theta dominance was quantified as the ratio of powers in the 5 to 10 Hz and 2 to 16 Hz frequency bands. EMG was estimated as the zero-lag correlation be- tween filtered (300 to 600 Hz) signals across recording sites (55). Soft sticky thresholds on these metrics were used to identify states. High LFP principal component 1 and the low EMG were considered non-REM, the high theta and low EMG were considered REM, and the remain- ing data were taken to reflect the waking state. For analysis of neural activity during active behavior (e.g., place cell and theta sequence analysis), only periods in which the animals run faster than 5 cm/s were included. SWR detection To detect SWRs, one channel around the CA1 pyramidal layer was chosen for the ripple de- tection, and one channel from CA1 stratum radiatum was chosen for sharp wave detection. The difference between the two channels was used as the basis for SWR detection. This dif- ference signal was filtered with a low-pass filter at 55 Hz, and then local minima were detected as candidate events. The corresponding ripple power (amplitude of the difference signal in the bandpass 80 to 250 Hz) for each candidate event was recorded. A true event was considered when both a sharp wave and ripple oscillation were detected in the same window. K-means clustering was used to define clustering of SWR from non-SWR events, and manual curation was used to better define the boundary and remove outliers. The events were then expanded until the (nonclipped) ripple power fell below 1 SD; short events (<15 ms) and longer events (>400 ms) were discarded. To refine the detection of SWR start and end points for reactivation and replay analyses (in order to avoid including empty bins), we detected population burst events as periods when the instantaneous population firing rate (binned in 1-ms bins and smoothed with a Gaussian kernel with 10-ms width) reached a peak of >2 SD and remained greater than the mean. The SWR start and end time points were set as the start and end time points of the population burst within the respective SWR, and SWRs without population bursts were discarded. Place cell analysis Spiking data and the tracked animal’s posi- tion were binned into 3-cm-wide segments of Liu et al., Science 382, eadi8237 (2023) 20 October 2023 10 of 16 RES EARCH | R E S E A R C H A R T I C L E the camera field projected onto the maze floor, generating raw maps of spike counts and oc- cupancy. A Gaussian kernel (SD = 3 cm) was applied to both raw maps of spike and oc- cupancy, and a smoothed rate map was con- structed by dividing the spike map by the occupancy map. Independent rate maps were constructed for the different running direc- tions in the mazes. Only periods in which the animal velocity was >4 cm/s were included. A place field was defined as a continuous re- gion of at least 15 cm2, where the mean firing rate was >10% of the peak rate in the maze, and the peak firing rate was >3Hz. Spatial information (SI) for individual place cells was obtained from the linearized rate maps (97) SI ¼ XN i¼1 pi li l log2 li l where N is the total number of spatial bins, pi is the probability of occupancy, and li is the firing rate in the ith spatial bin, and l the average firing rate of the cell. To measure the experience-dependent change of SI across laps in each condition, the mean and standard de- viation of SI across all cells within a condition group were calculated, and the SI was then z-scored for that condition (Fig. 3B). To identify cells that discriminate between the two sides of the CPP task, we constructed a vector for each cell containing the mean firing rate for every movement interval that the ani- mal spent in each of the two halves. We then performed a one-tailed Wilcoxon rank sum test between the firing vectors in the two halves, and cells for which this comparison was sig- nificant were identified as cells with firing specific to that respective half. To quantify discrimination in post-task sleep, pairwise post- task sleep correlations (CPOST, see below) were compared between pairs of cells specific for the same half versus pairs of cells that were specific for opposite halves of the environment. Population vector analysis To estimate the stability of the population code in the Stim OFF and Stim ON conditions, we performed population vector analysis as described previously (98). Briefly, we measured the correlation between the population firing curves in even and in odd trials for each run- ning direction, and we compared this value with the value obtained in surrogate data, in which cell identity was shuffled across even and odd trials. In addition, we performed a cross-correlation between the population firing curves in even and odd trials separately for each pair of spa- tial bins (fig. S3, A and B). To estimate rep- resentation stability, we computed the “error” defined as the distance between the peak correlation for each spatial bin (each of the columns of the matrix) to the diagonal. To statistically test the directionality of the hippo- campal code, we compared the representation stability within a given direction (even versus odd trajectories in the same direction) to the stability of the representation between the two directions of movement (even trajectories in one direction versus odd trajectories in the other direction). COM shift To measure PF shifting across laps (Fig. 3D), we used place cells that had PFs on the linear track, and calculated the center-of-mass (COM) of PFs for each lap n (COMn) (99) COMn ¼ X FRixiX FRi i i where FRi is the firing rate in the spatial bin i, and xi is the distance of the spatial bin i from the start of the running trajectory. For each lap, COM shift was then measured as the dif- ference between the first lap and the current lap. To prevent any edge effect, only PFs with a peak location outside the two ends of the track (i.e., defined by 20% of the track length) were included. For place cells with multiple PFs, only the primary PF (with the maximal peak rate) was included. Place field eccentricity To calculate the eccentricity of 2D place fields in the CPP task (fig. S12), we first defined the PF boundary on the 2D rate maps. PF was first defined as the area with firing rates larger than 40% of the peak rate. This area was fur- ther refined with a series of morphological operations: opening, closing, and infill, with Matlab functions imopen, imclose, and imfill, respectively. PF boundary was then detected on the resulting binarized image using the Matlab function bwboundaries. The eccentric- ity of the detected PF was measured using the Matlab function regionprops. For place cells with multiple PFs, only the primary PF (with the maximal peak rate located within the PF boundary with the maximal size) was included. The boundary and nonboundary cells were defined as the cells with the peak of the pri- mary PF located within and outside 15 cm of the wall, respectively. Phase precession Phase precession analysis was performed as previously described (34, 100). Circular-linear regression between relative position within the place field and theta phase was applied to calculate the phase-precession slope and correlation strength (101). The slope and cor- relation strength (r2) of phase precession were derived from this circular-linear regression analysis. Theta compression Theta compression analysis was performed as described previously (102), independently for the Stim OFF and the Stim ON direction. Briefly, for each pair of overlapping place cells, we computed (i) the distance between their place field peaks and (ii) the theta time lag. To qualify pairs with a significant theta time lag, we computed a coarse cross-correlogram for each cell pair using for durations T1 s using 5-ms bins. Cell pairs with a peak in this coarse correlogram within ±100 ms with at least five spikes in the peak bin were deemed signifi- cant, and other cell pairs were excluded from this analysis. To compute the theta time lag for the selected cell pairs, cross-correlograms were restricted to in-field spikes during run- ning periods of the respective direction (Stim OFF versus Stim ON) using 1-ms bins, and each cross-correlogram was filtered with a bandpass filter between 2 and 30 Hz. The time of the cross-correlogram peak was defined as the theta time lag of the pair of place cells. Theta compression slopes were computed by performing linear regression between the dis- tance between the place field peaks and the theta time lags. Linearizing positions on the cheeseboard maze Positions on the cheeseboard maze were lin- earized along an optimal trajectory, connecting the start box with the three rewarded loca- tions for that session (with the closest reward locations connecting directly to the start box). The animal’s linearized position was defined as the relative position along the optimal trajectory closest to the animal’s current position. To decode animal position from neu- ronal activity (for theta sequence and replay analyses), we only retained periods when this linearized position described the current ani- mal’s position well, defined as periods in which (i) the error distance between the opti- mal trajectory and the actual animal position was within 20 cm and (ii) the animal running speed along the optimal trajectory was at least 10 cm/s. To decode replay of familiar routes to previously learned trajectories (fig. S7), we modified the above procedure to de- fine the linearized trajectory: Instead of using the three currently rewarded locations, we used the three rewarded locations from the previ- ous day’s session. Decoding animal position from neuronal activity To decode position from place cell activity during both theta states and replay, we used a Bayesian reconstruction approach as described previously (65). Briefly, we computed the aver- ON(x) age firing rate probability li for each pyramidal cell i at position x as the normalized firing rate curve (spatial bin size: 1.5 cm) of the cell during running epochs in the OFF and the ON direction, respectively, OFF(x) and li Liu et al., Science 382, eadi8237 (2023) 20 October 2023 11 of 16 RES EARCH | R E S E A R C H A R T I C L E which constituted the training step in our decoder. To then decode the animal posi- tion from the neuronal activity as expressed by the firing rate vector n in a particular window of width ts, we estimate the proba- bility P(x|n) P xjnð Þ ¼ P njxð Þ (cid:3) P xð Þ P nð Þ where spikes are assumed to fire as indepen- dent Poisson processes P njxð Þ ¼ YN i¼1 ni Þ ð li xð Þ (cid:3) t n! i e(cid:2)li xð Þ(cid:3)t We applied the same procedure to decode linearized positions on the cheeseboard maze, using periods of movement along the line- arized trajectory in which the animal stayed within 20 cm of the optimal trajectory to train the decoder. Theta sequences Theta reconstruction matrices To decode the animal position during theta cycles, we trained Bayesian decoders for the OFF and the ON directions as described above and decoded the animal position during run- ning periods in the OFF and the ON direction, respectively, using t = 20 ms bins with 5-ms sliding window. We then computed theta re- construction matrices for each theta cycle while the animal was running in the OFF and the ON direction by dividing the cycle in 500 temporal bins and interpolating the respective (OFF or ON) decoded position. The same procedure was applied to construct theta reconstruction matrices for Stim OFF and Stim ON cheese- board sessions using the linearized cheeseboard position during periods of movement along the linearized trajectory and stayed within 20 cm of the optimal trajectory. Theta cycles To detect theta cycles, we first detected deep and superficial CA1 sublayers using the depth profile of SWRs as described previously (51) and selected the deep CA1 channel with the highest theta power in the normalized spec- trum. Theta cycles were defined as the peaks in the LFP filtered in the theta band during running periods, and cycles shorter than 100 ms or longer than 200 ms were excluded from further analyses. To correct for the phase shift that can occur in different sublayer depth of the theta channel, the final theta cycles were shifted to ensure that the theta peaks were de- fined as the phase of maximal uncertainty of the theta reconstruction matrices, where un- certainty u was defined as u ¼ (cid:2)maxx g. Theta peaks correspond 0° (0 rad.) and 360° (2p rad.) and troughs at 180° (p rad.) and 540° (3p rad.) of theta waves. P xjnð f Þ Theta sequences To quantify theta sequence, we applied the measures of quadrant scores and weighted correlation on the theta reconstruction mat- rices for each cycle. To compute the quadrant score, we divided the central region in the theta reconstruction matrix that was within 50 cm of the animal’s current location and within 1 2 p of the theta trough (p) into four equal quadrants, and computed the ratio be- tween the probability within quadrants II and IV (representing directions consistent with the animal’s running direction) and the prob- ability within quadrants I and III (represent- ing directions incongruent with the animal’s running direction). The weighted correlation measured the correlation between time t and location l weighted by the decoded probability values p within the reconstruction matrix w ¼ p cov t; ljp Þ ð ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Þcov l; ljp cov t; tjp Þ ð ð where cov t; ljp ð m tjpð Þ ¼ X Þ ¼ X pitiX pi i i pi ti (cid:2) m tjpð X ð i Þ li (cid:2) m ljpð Þ ð pi Þ Þ , i , and m ljpð Þ ¼ X piliX pi i . i The look-ahead index of theta sequences is calculated as in previous reports (46). In brief, we recomputed the reconstruction matrix quad- rants, this time extending the reach to use all decoded probabilities within 1 m of the ani- mal’s current location. The look-ahead index was measured by comparing the probabilities in the quadrants ahead (quadrant IV, future) and behind (quadrant I, past) in the second half of theta phases, as (IV – I)/(IV + I). Replay analysis Candidate replay events We first detected periods of elevated pyra- midal cell population activity as defining periods when the instantaneous firing rate (1-ms bins, smoothed with a Gaussian kernel of width 10 ms) was greater than the mean and reached a peak of >2 SD above the mean during the session. Candidate events were between 100 and 500 ms in duration, and they each over- lapped with an independently detected SWR; otherwise they were excluded from further analyses. Quantifying replay We decoded the animal position for the OFF and the ON direction during each candidate replay event using non-overlapping t = 20 ms bins (but see fig. S7A for replay results with t = 40 ms and t = 60 ms). OFF and ON direction replay were scored independently. To assess replay quality, we computed trajectory scores of the reconstruction matrices (103). Briefly, each reconstruction matrix was fitted with a line (of slope a and intercept b) maximizing the average likelihood R that the animal is located within a distance d (set to 22.5 cm) of the linear trajectory defined by the slope a and intercept b R v; rð Þ ¼ 1 m Xm(cid:2)1 k¼0 P jpos (cid:2) r (cid:2) v (cid:3) k (cid:3) Dxj≤ d ð ð Þ Þ The score of each event was normalized by subtracting the mean and dividing by the stan- dard deviation of a distribution of 500 shuf- fled scores obtained by independently shifting each of the columns of the event’s reconstruc- tion matrix along the spatial dimension by a random distance. An event with a score ex- ceeding the scores of 95% of its shuffled scores was considered significant. To quantify replay, we normalized the proportion of significant events during awake behavior or post-task sleep by subtracting the mean proportion of significant events in pre-task sleep. In a control analysis, we corrected for the differences in decoding quality between Stim OFF and Stim ON trajectories (fig. S7C). For this purpose, downsampled units before decod- ing the Stim OFF trajectories by progressively dropping units until the average decoded error in Stim OFF trajectories was larger than the average decoded error in Stim ON trajec- tories, and we recomputed replay events for Stim OFF trajectories using this more limited sample (fig. S7C). We performed an additional quantification of replay, where for each replay event we com- puted the weighted correlation and the jump distance. The weighted correlation measures the correlation coefficient between time and decoded space, and jump distance was defined as the maximum distance between the peak decoded positions in two successive time bins of the same event. Trajectory events were defined as the events with high weighted correlation (>0.6) and low jump distance (<75 cm), and to assess whether the number of trajectory events exceeded chance, we compared the number trajectory events in the recorded data with the number of trajectory events in surrogate data where cell identities were shuffled 500 times. This comparison yielded a P value (the propor- tion of shuffled datasets with as many or more trajectory events than those observed in the original dataset), and we generated a signifi- cance matrix of P values for progressively stricter weighted correlation thresholds (0.6, 0.65, 0.7, 0.75, 0.8, 0.85, 0.9, 0.95) and jump distance thresholds (75 cm, 62.5 cm, 50 cm, 37.5 cm, 25 cm, 12.5 cm). Pairwise bias correlations The above replay analyses test for the replay of stereotypical trajectories and are not directly applicable to the trajectory-free behavior in the Liu et al., Science 382, eadi8237 (2023) 20 October 2023 12 of 16 RES EARCH | R E S E A R C H A R T I C L E CPP task. To assess whether the order of unit spiking in behavior is preserved in these se- ssions, we computed the similarity between behavioral sequences and SWR sequences using a pairwise bias correlation method (65). Briefly, we computed a bias matrix Bk for each sequence k Bk ¼ nk i; j ð nk ið Þ (cid:3) nk jð Þ Þ where nk(i) is the number of spikes emitted by neuron i in the kth sequence, and nk(i, j) is the number of times neuron i spiked before neu- ron j in the kth sequence (see fig. S12F for an example). Bk(i, j) therefore reflects the bias of i to spike before j in the kth sequence, taking values between 0 (i never precedes j) and 1 (i always precedes j), with 0.5 representing no bias (i precedes j half of the time). The cor- relation between two sequences can be esti- mated as the cosine between their skew-bias matrices as each bias matrix is unwrapped into a vector 2Bk – 2 (this normalization is to achieve a range of values between −1 and 1, assuring that two mirror sequences would have opposite signs with vectors pointing in opposite directions). The activity of cells that are not common to the two sequences is ignored. To test whether the order of firing in be- havior was preserved in post-task sleep, we computed the mean bias matrix of all se- quences taking place in theta cycles during the task in a given recording session. We com- puted the correlation (cosine) between this bias matrix and each of the bias matrices de- scribing sequences in post-task sleep SWRs. We compared this correlation with a shuffled distribution obtained by recomputing the cor- relation after shuffling the spikes sequence n = 100 times and deemed as significant those se- quences with correlations whose absolute value exceeded 5% of the shuffled distribution. We normalized the proportion of significant events by subtracting the mean proportion of signifi- cant events in pre-task sleep. Coactivity and reactivation analysis Cofiring Each spike train was binned in individual theta cycles, and the cofiring of a pair of cells was defined as the Pearson correlation coefficient of the binned firing rates of the two cells across all theta cycles in the recorded session. The matrix of Pearson correlation coefficients was called the cofiring matrix C. Explained variance Explained variance (EV) was quantified as pre- viously described (104). Briefly, for each session, we computed the proportion of the variance in the population spiking activity in post-task sleep that is explained by task-related activity, after taking into account correlations pre- existing in sleep before the experience 0 B B @ r EV ¼ 1 2 rC;CPOST (cid:2) rC;CPRE ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (cid:7) (cid:6) (cid:6) 1 (cid:2) r2 1 (cid:2) r2 rCPOST;CPRE (cid:7) CPOST;CPRE C;CPRE C C A where C, CPRE, and CPOST refer to the cofiring matrix and the correlation matrices across SWRs in pre- and post-task sleep, respectively, and rA,B describes the correlation coefficient between the pairwise correlation matrices A and B. EV was computed for each recorded session separately, and it was compared with the control value of reverse explained variance (REV) obtained by recomputing EV after swap- ping CPRE and CPOST. Pairwise reactivation For the pairwise reactivation analyses, we used linear regression between CPOST and CPRE to remove the portion of the variance in CPOST that can be explained by preexisting pat- terns, and we correlated the resultant resid- uals CRESIDUAL with C (105). For the CPP reactivation analysis, we iden- tified cells that had a preference for each of the two sides of the maze by computing each cell’s firing rate for every movement period that the animal spent moving in each of the two sides of the maze. Cells with a significant preference (Wilcoxon rank sum test, alpha a = 0.05) were divided according to the sign of their prefer- ence between rewarded-side responsive cells and unrewarded-side responsive cells. We then compared CRESIDUAL (see above) in pairs of rewarded-side responsive cells versus pairs of unrewarded-side responsive cells. Cell assembly analyses For detecting cell assembly patterns, we used an unsupervised statistical framework based on a hybrid principal components analysis (PCA) followed by independent component analysis (ICA) as previously reported (49, 106–108). In brief, spike trains of each neuron were binned in 25-ms intervals for the whole session (in- cluding only task periods during movement), the matrix of firing correlation coefficients for all pairs of neurons were constructed. Next, we calculated the number of assemblies on the basis of those principal components whose eigenvalues exceeded the threshold for ran- dom firing correlations (using the Marchenko- Pastur law). This method provides a number of significant patterns smaller than the num- ber of neurons. Each of these patterns explained more variance of the spike train correlation ma- trix than other patterns that would result from independently firing neurons. Independent component analysis (fast-ICA algorithm) was then used to determine for each assembly (component) the vector of weights with which each neuron’s firing contributes to that as- sembly. The strength of each assembly i’s ac- tivation for a given time bin k was computed as follows Sik ¼ Z T k PZk where Z is the population activity matrix of the z-scored firing rate of each unit, and P is the outer product of the component i’s weights, in which the diagonal has been set to zero so that isolated spikes from individual units do not contribute to S. Most of the detected assembly patterns con- sisted of a few neurons with high weights and a large group of neurons with weights around zero. Assembly members were thus consi- dered those cells whose weight exceeded the mean weight of the assembly by two standard deviations (107). To produce the assembly activity profiles in fig. S9B, we took the assembly activation strength during the last hour of cumulative baseline sleep and the first hour of cumulative post- task sleep, excluding periods outside of non- REM sleep. During the task, we took the assembly activation strength during running in Stim OFF and Stim ON trajectories sepa- rately. Because the exploration time is variable between sessions, instead of absolute elapsed time, we divided the Stim OFF and Stim ON exploration time into quantiles. Assembly activation events were detected as peaks in the assembly activation strength ex- ceeding a value of 5 (109). The point process of assembly activation events during behavior in the task was used to compute the spatial infor- mation content of assembly activity in fig. S4G. The order of assembly activation events in SWRs was analyzed in fig. S9E as follows. For each SWR in which at least three assemblies were active, we correlated the timing of assem- bly activations within an event to the respec- tive positions (that is, the location of peak activity for a given assembly). We took the absolute value of the Spearman’s rank-order correlation coefficient r to account for reverse replay. To correct for coefficient variation with the number of assemblies (e.g., fewer points yield higher correlation coefficients) and the day of the recording, we subtracted from each n defined as the coefficient r a baseline value r0 mean r of SWRs recorded pre-task sleep of the same session with a matching number n of assembly activations. Spectral analysis, cross-frequency coupling, and spike-LFP coupling To obtain the phase of the theta rhythm, one LFP channel was selected as explained above. LFP was bandpass-filtered in the range of 5 to 15 Hz. Theta phase was then computed using the Hilbert transform of the filtered LFP. Liu et al., Science 382, eadi8237 (2023) 20 October 2023 13 of 16 RES EARCH | R E S E A R C H A R T I C L E Cross-frequency coupling To perform spectral analysis at a high resolu- tion in time and frequency, the complex wave- let transform (CWT) of the LFP (or ICs) was calculated using complex Morlet wavelets. Wavelets were calculated using a logarith- mically spaced frequency vector in the band of interest (25 to 200 Hz). Phase-amplitude cross-frequency coupling for a given LFP re- cording was assessed using the modulation index measure (MI) (110). Phase time-series were binned into phase intervals and the mean wavelet amplitude was calculated for each of them. The MI was obtained by measuring the divergence of the observed amplitude distribu- tion from the uniform distribution. The sta- tistical significance of the MI values (P value) was assessed by a surrogate analysis (n = 1000 surrogates) with random shifts between the phase and amplitude time series. For the pres- ented plots, grand averages were calculated as the mean across all animals, unless otherwise indicated, and MI reported were significant [P < 0.01 compared with surrogate distribu- tion (34, 55)]. Spike-LFP coupling The phase-locking of spikes to LFP features at each frequency was measured for individual units using the wavelet phase from 25 to 200 Hz (30 logarithmically spaced wavelet scales) at the time of each spike (34, 55). Only neurons that fired at least 100 spikes during the se- lected task intervals were included in the analy- sis. Reference LFP was taken 200 to 400 mm away from the electrode, where the unit was recorded to minimize spike energy leakage into the LFP. Modulation indices were calculated using the mean resultant length of the phases, and significance was estimated using the Ray- leigh test for nonuniformity (P < 0.05) using circular statistics. Preferred frequency of mod- ulation was determined as the largest mean vector length of each significantly modulated neuron. The mean angle of the phases for a given neuron’s spikes was taken as the pref- erred phase. Independent component and current source density analysis of LFPs To separate the different sources that con- tribute to the LFP mixed signal, we used an approach based on independent component analysis (ICA) that has been described and validated previously for hippocampal record- ings (41, 55, 111). Here, we applied ICA to spatially contiguous LFP channels after filtering in the gamma band (25 to 200 Hz). The ICA algorithm (runinca) (112) takes a time series of data with dimension equal to the number of recording sites and returns a time series of the same dimensionality, but rotated so that each dimension represents a different IC. The inverse of the mixing matrix that transforms the LFP data into the ICs gives the channel weight of each component that is captured for each recording site. When projected back to the anatomical location of the recording site, this corresponds to the spatial voltage loadings of each IC (41, 55, 111). Once ICs have been extracted from the raw LFP traces, they can be analyzed as if they were active independently from activities at other locations. Hierarchical bootstrap Hierarchical bootstrap (113) was performed to analyze data with hierarchical structure. Briefly, bootstrap datasets were created by resampling with replacement following levels of hierar- chical order (in the order of animals followed by sessions). The mean of each resampled boot- strap data was calculated each time, in a total of 1000 resampling times. The final statistics were done on the populations of resampled data from the different experimental condi- tions. P value indicated the test result that the values of one condition were higher than the values of the other condition after controlling for the other nesting variables. Statistical analyses Statistical analyses were performed with MATLAB functions or custom-made scripts. No specific analysis was used to estimate minimal popu- lation sample or group size, but the number of animals, sessions, and recorded cells was larger or similar to those used in previous related works (41, 65, 66, 75, 92). The unit of analysis was typically identified as single neurons or assemblies. In a few cases, the unit of analysis was sessions or animals, and this is stated in the text. Unless otherwise noted, nonparametric two-tailed Wilcoxon rank sum (equivalent to Mann-Whitney U test) or Wilcoxon signed- rank test was used for unpaired and paired data, respectively. For multiple comparisons following analysis of variance (ANOVA), Tukey’s honest significant difference post-hoc test was used. All statistical tests were two-sided, unless otherwise specified. Throughout the manuscript, statistical significance is indicated by asterisks: *P < 0.05, **P < 0.01, ***P < 0.001. On box plots, the central mark indicates the median; bottom and top edges of the box indicate the 25th and 75th percentiles, respec- tively; and whiskers extend to the most extreme data points not considered outliers. Outliers are not displayed in some plots but were included in statistical analysis. Owing to experimental design constraints, the experimenter was not blind to the manipulation performed during the experiment (i.e., optogenetic manipulation). RE FERENCES AND NOTES 7. 5. 2. 3. 4. 6. P. Tovote, J. P. Fadok, A. Lüthi, Neuronal circuits for fear and anxiety. Nat. Rev. Neurosci. 16, 317–331 (2015). doi: 10.1038/nrn3945; pmid: 25991441 G. R. I. Barker, E. C. Warburton, When is the hippocampus involved in recognition memory? J. Neurosci. 31, 10721–10731 (2011). doi: 10.1523/JNEUROSCI.6413-10.2011; pmid: 21775615 A. V. Beylin et al., The role of the hippocampus in trace conditioning: Temporal discontinuity or task difficulty? Neurobiol. Learn. Mem. 76, 447–461 (2001). doi: 10.1006/ nlme.2001.4039; pmid: 11726247 C. J. MacDonald, K. Q. Lepage, U. T. Eden, H. Eichenbaum, Hippocampal “time cells” bridge the gap in memory for discontiguous events. Neuron 71, 737–749 (2011). doi: 10.1016/j.neuron.2011.07.012; pmid: 21867888 E. C. Tolman, Cognitive maps in rats and men. Psychol. Rev. 55, 189–208 (1948). doi: 10.1037/h0061626; pmid: 18870876 E. A. Maguire, K. Woollett, H. J. Spiers, London taxi drivers and bus drivers: A structural MRI and neuropsychological analysis. Hippocampus 16, 1091–1101 (2006). doi: 10.1002/ hipo.20233; pmid: 17024677 D. Dupret, J. O’Neill, B. Pleydell-Bouverie, J. Csicsvari, The reorganization and reactivation of hippocampal maps predict spatial memory performance. Nat. Neurosci. 13, 995–1002 (2010). doi: 10.1038/nn.2599; pmid: 20639874 H. C. Barron et al., Neuronal computation underlying inferential reasoning in humans and mice. Cell 183, 228–243.e21 (2020). doi: 10.1016/j.cell.2020.08.035; pmid: 32946810 10. D. Hassabis, D. Kumaran, S. D. Vann, E. A. Maguire, Patients with hippocampal amnesia cannot imagine new experiences. Proc. Natl. Acad. Sci. U.S.A. 104, 1726–1731 (2007). doi: 10.1073/pnas.0610561104; pmid: 17229836 J. Widloski, D. J. Foster, Flexible rerouting of hippocampal replay sequences around changing barriers in the absence of global place field remapping. Neuron 110, 1547–1558.e8 (2022). doi: 10.1016/j.neuron.2022.02.002; pmid: 35180390 J. O’Keefe, J. Dostrovsky, The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Res. 34, 171–175 (1971). doi: 10.1016/0006-8993(71) 90358-1; pmid: 5124915 12. 11. 8. 9. 13. H. Eichenbaum, P. Dudchenko, E. Wood, M. Shapiro, H. Tanila, The hippocampus, memory, and place cells: Is it spatial memory or a memory space? Neuron 23, 209–226 (1999). doi: 10.1016/ S0896-6273(00)80773-4; pmid: 10399928 F. P. Battaglia, K. Benchenane, A. Sirota, C. M. A. Pennartz, S. I. Wiener, The hippocampus: Hub of brain network communication for memory. Trends Cogn. Sci. 15, 310–318 (2011). pmid: 21696996 S. A. Josselyn, S. Köhler, P. W. Frankland, Finding the engram. Nat. Rev. Neurosci. 16, 521–534 (2015). doi: 10.1038/nrn4000; pmid: 26289572 R. L. Buckner, The role of the hippocampus in prediction and imagination. Annu. Rev. Psychol. 61, 27–48 (2010). doi: 10.1146/annurev.psych.60.110707.163508; pmid: 19958178 K. L. Stachenfeld, M. M. Botvinick, S. J. Gershman, The hippocampus as a predictive map. Nat. Neurosci. 20, 1643–1653 (2017). doi: 10.1038/nn.4650; pmid: 28967910 14. 15. 16. 17. 18. H. Eichenbaum, N. J. Cohen, Can we reconcile the declarative memory and spatial navigation views on hippocampal function? Neuron 83, 764–770 (2014). doi: 10.1016/ j.neuron.2014.07.032; pmid: 25144874 A. M. Wikenheiser, A. D. Redish, Decoding the cognitive map: Ensemble hippocampal sequences and decision making. Curr. Opin. Neurobiol. 32, 8–15 (2015). doi: 10.1016/ j.conb.2014.10.002; pmid: 25463559 19. 20. D. O. Hebb, The Organization of Behavior: A Neuropsychological 21. Theory (Wiley, 1949). K. D. Harris, J. Csicsvari, H. Hirase, G. Dragoi, G. Buzsáki, Organization of cell assemblies in the hippocampus. Nature 424, 552–556 (2003). doi: 10.1038/nature01834; pmid: 12891358 22. M. El-Gaby et al., An emergent neural coactivity code for dynamic memory. Nat. Neurosci. 24, 694–704 (2021). doi: 10.1038/s41593-021-00820-w; pmid: 33782620 23. W. E. Skaggs, B. L. McNaughton, M. A. Wilson, C. A. Barnes, Theta phase precession in hippocampal neuronal populations and the compression of temporal sequences. Hippocampus 6, 149–172 (1996). doi: 10.1002/(SICI)1098-1063(1996) 6:2<149::AID-HIPO6>3.0.CO;2-K; pmid: 8797016 24. G. Dragoi, G. Buzsáki, Temporal encoding of place 1. K. Nakazawa et al., Requirement for hippocampal CA3 NMDA receptors in associative memory recall. Science 297, 211–218 (2002). doi: 10.1126/science.1071795; pmid: 12040087 sequences by hippocampal cell assemblies. Neuron 50, 145–157 (2006). doi: 10.1016/j.neuron.2006.02.023; pmid: 16600862 Liu et al., Science 382, eadi8237 (2023) 20 October 2023 14 of 16 RES EARCH | R E S E A R C H A R T I C L E 25. D. J. Foster, M. A. Wilson, Hippocampal theta sequences. Hippocampus 17, 1093–1099 (2007). doi: 10.1002/ hipo.20345; pmid: 17663452 26. W. E. Skaggs, B. L. McNaughton, Replay of neuronal firing sequences in rat hippocampus during sleep following spatial experience. Science 271, 1870–1873 (1996). doi: 10.1126/ science.271.5257.1870; pmid: 8596957 48. 27. A. K. Lee, M. A. Wilson, Memory of sequential experience in the hippocampus during slow wave sleep. Neuron 36, 1183–1194 (2002). doi: 10.1016/S0896-6273(02)01096-6; pmid: 12495631 28. G. Girardeau, K. Benchenane, S. I. Wiener, G. Buzsáki, M. B. Zugaro, Selective suppression of hippocampal ripples impairs spatial memory. Nat. Neurosci. 12, 1222–1223 (2009). doi: 10.1038/nn.2384; pmid: 19749750 29. S. P. Jadhav, C. Kemere, P. W. German, L. M. Frank, Awake 30. 31. 32. hippocampal sharp-wave ripples support spatial memory. Science 336, 1454–1458 (2012). doi: 10.1126/ science.1217230; pmid: 22555434 I. Gridchyn, P. Schoenenberger, J. O’Neill, J. Csicsvari, Assembly-specific disruption of hippocampal replay leads to selective memory deficit. Neuron 106, 291–300.e6 (2020). doi: 10.1016/j.neuron.2020.01.021; pmid: 32070475 Y. Wang, S. Romani, B. Lustig, A. Leonardo, E. Pastalkova, Theta sequences are essential for internally generated hippocampal firing fields. Nat. Neurosci. 18, 282–288 (2015). doi: 10.1038/nn.3904; pmid: 25531571 T. Nakashiba, D. L. Buhl, T. J. McHugh, S. Tonegawa, Hippocampal CA3 output is crucial for ripple-associated reactivation and consolidation of memory. Neuron 62, 781–787 (2009). doi: 10.1016/j.neuron.2009.05.013; pmid: 19555647 33. M. I. Schlesiger et al., The medial entorhinal cortex is necessary for temporal organization of hippocampal neuronal activity. Nat. Neurosci. 18, 1123–1132 (2015). doi: 10.1038/ nn.4056; pmid: 26120964 55. 34. A. Fernández-Ruiz et al., Entorhinal-CA3 dual-input control of spike timing in the hippocampus by theta-gamma coupling. Neuron 93, 1213–1226.e5 (2017). doi: 10.1016/ j.neuron.2017.02.017; pmid: 28279355 36. 35. B. Lasztóczi, T. Klausberger, Hippocampal place cells couple to three different gamma oscillations during place field traversal. Neuron 91, 34–40 (2016). doi: 10.1016/ j.neuron.2016.05.036; pmid: 27387648 I. Zutshi, M. Valero, A. Fernández-Ruiz, G. Buzsáki, Extrinsic control and intrinsic computation in the hippocampal CA1 circuit. Neuron 110, 658–673.e5 (2022). doi: 10.1016/ j.neuron.2021.11.015; pmid: 34890566 J. B. Hales et al., Medial entorhinal cortex lesions only partially disrupt hippocampal place cells and hippocampus- dependent place memory. Cell Rep. 9, 893–901 (2014). doi: 10.1016/j.celrep.2014.10.009; pmid: 25437546 37. 39. 40. 41. 42. learning-related changes in CA1 representations. Nature 611, 554–562 (2022). doi: 10.1038/s41586-022-05378-6; pmid: 36323779 J. Yamamoto, S. Tonegawa, Direct medial entorhinal cortex input to hippocampal CA1 is crucial for extended quiet awake replay. Neuron 96, 217–227.e4 (2017). doi: 10.1016/ j.neuron.2017.09.017; pmid: 28957670 J. Dimidschstein et al., A viral strategy for targeting and manipulating interneurons across vertebrate species. Nat. Neurosci. 19, 1743–1749 (2016). doi: 10.1038/nn.4430; pmid: 27798629 A. Fernández-Ruiz et al., Gamma rhythm communication between entorhinal cortex and dentate gyrus neuronal assemblies. Science 372, eabf3119 (2021). doi: 10.1126/ science.abf3119; pmid: 33795429 J. O’Keefe, M. L. Recce, Phase relationship between hippocampal place units and the EEG theta rhythm. Hippocampus 3, 317–330 (1993). doi: 10.1002/ hipo.450030307; pmid: 8353611 43. A. M. Wikenheiser, A. D. Redish, Hippocampal theta sequences reflect current goals. Nat. Neurosci. 18, 289–294 (2015). doi: 10.1038/nn.3909; pmid: 25559082 44. A. Johnson, A. D. Redish, Neural ensembles in CA3 transiently encode paths forward of the animal at a decision point. J. Neurosci. 27, 12176–12189 (2007). doi: 10.1523/ JNEUROSCI.3761-07.2007; pmid: 17989284 45. K. Kay et al., Constant sub-second cycling between representations of possible futures in the hippocampus. Cell 180, 552–567.e25 (2020). doi: 10.1016/ j.cell.2020.01.014; pmid: 32004462 46. W. Tang, J. D. Shin, S. P. Jadhav, Multiple time-scales of decision-making in the hippocampus and prefrontal cortex. eLife 10, e66227 (2021). doi: 10.7554/eLife.66227; pmid: 33683201 47. M. R. Mehta, C. A. Barnes, B. L. McNaughton, Experience- dependent, asymmetric expansion of hippocampal place fields. Proc. Natl. Acad. Sci. U.S.A. 94, 8918–8921 (1997). doi: 10.1073/pnas.94.16.8918; pmid: 9238078 L. Genzel et al., A consensus statement: Defining terms for reactivation analysis. Philos. Trans. R. Soc. London Ser. B 375, 20200001 (2020). doi: 10.1098/rstb.2020.0001; pmid: 32248790 49. A. Oliva, A. Fernández-Ruiz, F. Leroy, S. A. Siegelbaum, Hippocampal CA2 sharp-wave ripples reactivate and promote social memory. Nature 587, 264–269 (2020). doi: 10.1038/ s41586-020-2758-y; pmid: 32968277 51. 50. K. Diba, G. Buzsáki, Forward and reverse hippocampal place- cell sequences during ripples. Nat. Neurosci. 10, 1241–1242 (2007). doi: 10.1038/nn1961; pmid: 17828259 R. E. Harvey, H. L. Robinson, C. Liu, A. Oliva, A. Fernandez-Ruiz, Hippocampo-cortical circuits for selective memory encoding, routing, and replay. Neuron 111, 2076–2090.e9 (2023). doi: 10.1016/j.neuron.2023.04.015; pmid: 37196658 T. M. George, W. de Cothi, K. L. Stachenfeld, C. Barry, Rapid learning of predictive maps with STDP and theta phase precession. eLife 12, e80663 (2023). doi: 10.7554/ eLife.80663; pmid: 36927826 52. 53. A. Alvernhe, E. Save, B. Poucet, Local remapping of place cell firing in the Tolman detour task. Eur. J. Neurosci. 33, 1696–1705 (2011). doi: 10.1111/j.1460-9568.2011.07653.x; pmid: 21395871 54. S. Tanni, W. de Cothi, C. Barry, State transitions in the statistically stable place cell population correspond to rate of perceptual change. Curr. Biol. 32, 3505–3514.e7 (2022). doi: 10.1016/j.cub.2022.06.046; pmid: 35835121 E. W. Schomburg et al., Theta phase segregation of input- specific gamma patterns in entorhinal-hippocampal networks. Neuron 84, 470–485 (2014). doi: 10.1016/ j.neuron.2014.08.051; pmid: 25263753 56. V. Lopes-Dos-Santos et al., Parsing hippocampal theta 57. 58. oscillations by nested spectral components during spatial exploration and memory-guided behavior. Neuron 100, 940–952.e7 (2018). doi: 10.1016/j.neuron.2018.09.031; pmid: 30344040 L. L. Colgin et al., Frequency of gamma oscillations routes flow of information in the hippocampus. Nature 462, 353–357 (2009). doi: 10.1038/nature08573; pmid: 19924214 J. Csicsvari, H. Hirase, A. Mamiya, G. Buzsáki, Ensemble patterns of hippocampal CA3-CA1 neurons during sharp wave–associated population events. Neuron 28, 585–594 (2000). doi: 10.1016/S0896-6273(00)00135-5; pmid: 11144366 temporal coding in the CA1 ensemble. Nat. Neurosci. 19, 945–951 (2016). doi: 10.1038/nn.4311; pmid: 27239937 60. H. Davoudi, D. J. Foster, Acute silencing of hippocampal CA3 61. reveals a dominant role in place field responses. Nat. Neurosci. 22, 337–342 (2019). doi: 10.1038/s41593-018- 0321-z; pmid: 30664772 A. Ecker et al., Hippocampal sharp wave-ripples and the associated sequence replay emerge from structured synaptic interactions in a network model of area CA3. eLife 11, e71850 (2022). doi: 10.7554/eLife.71850; pmid: 35040779 62. G. Q. Bi, M. M. Poo, Synaptic modifications in cultured hippocampal neurons: Dependence on spike timing, synaptic strength, and postsynaptic cell type. J. Neurosci. 18, 10464–10472 (1998). doi: 10.1523/JNEUROSCI.18-24- 10464.1998; pmid: 9852584 67. A. Chenani et al., Hippocampal CA1 replay becomes less prominent but more rigid without inputs from medial entorhinal cortex. Nat. Commun. 10, 1341 (2019). doi: 10.1038/s41467-019-09280-0; pmid: 30902981 68. D. S. Olton, B. C. Papas, Spatial memory and hippocampal 69. 70. 71. function. Neuropsychologia 17, 669–682 (1979). doi: 10.1016/ 0028-3932(79)90042-3; pmid: 522981 E. R. Wood, P. A. Dudchenko, H. Eichenbaum, The global record of memory in hippocampal neuronal activity. Nature 397, 613–616 (1999). doi: 10.1038/17605; pmid: 10050854 J. O’Keefe, L. Nadel, The Hippocampus as a Cognitive Map (Oxford Univ. Press, 1978). T. E. J. Behrens et al., What is a cognitive map? Organizing knowledge for flexible behavior. Neuron 100, 490–509 (2018). doi: 10.1016/j.neuron.2018.10.002; pmid: 30359611 72. A. S. Gupta, M. A. A. van der Meer, D. S. Touretzky, A. D. Redish, Hippocampal replay is not a simple function of experience. Neuron 65, 695–705 (2010). doi: 10.1016/ j.neuron.2010.01.034; pmid: 20223204 73. H. F. Ólafsdóttir, C. Barry, A. B. Saleem, D. Hassabis, H. J. Spiers, Hippocampal place cells construct reward related sequences through unexplored space. eLife 4, e06063 (2015). doi: 10.7554/eLife.06063; pmid: 26112828 74. H. Igata, Y. Ikegaya, T. Sasaki, Prioritized experience replays on a hippocampal predictive map for learning. Proc. Natl. Acad. Sci. U.S.A. 118, e2011266118 (2021). doi: 10.1073/ pnas.2011266118; pmid: 33443144 J. D. Shin, W. Tang, S. P. Jadhav, Dynamics of awake hippocampal-prefrontal replay for spatial learning and memory-guided decision making. Neuron 104, 1110–1125.e7 (2019). doi: 10.1016/j.neuron.2019.09.012; pmid: 31677957 75. 76. D. Bendor, M. A. Wilson, Biasing the content of hippocampal replay during sleep. Nat. Neurosci. 15, 1439–1444 (2012). doi: 10.1038/nn.3203; pmid: 22941111 77. A. A. Carey, Y. Tanaka, M. A. A. van der Meer, Reward revaluation biases hippocampal replay content away from the preferred outcome. Nat. Neurosci. 22, 1450–1459 (2019). doi: 10.1038/s41593-019-0464-6; pmid: 31427771 78. A. K. Gillespie et al., Hippocampal replay reflects specific past experiences rather than a plan for subsequent choice. Neuron 109, 3149–3163.e6 (2021). doi: 10.1016/ j.neuron.2021.07.029; pmid: 34450026 79. K. C. Bittner, A. D. Milstein, C. Grienberger, S. Romani, 80. 81. J. C. Magee, Behavioral time scale synaptic plasticity underlies CA1 place fields. Science 357, 1033–1036 (2017). doi: 10.1126/science.aan3846; pmid: 28883072 L. Roux, B. Hu, R. Eichler, E. Stark, G. Buzsáki, Sharp wave ripples during learning stabilize the hippocampal spatial map. Nat. Neurosci. 20, 845–853 (2017). doi: 10.1038/nn.4543; pmid: 28394323 K. Liu, J. Sibille, G. Dragoi, Preconfigured patterns are the primary driver of offline multi-neuronal sequence replay. Hippocampus 29, 275–283 (2019). doi: 10.1002/hipo.23034; pmid: 30260526 82. A. D. Grosmark, G. Buzsáki, Diversity in neural firing dynamics supports both rigid and learned hippocampal sequences. Science 351, 1440–1443 (2016). doi: 10.1126/ science.aad1935; pmid: 27013730 83. A. Noguchi, R. Huszár, S. Morikawa, G. Buzsáki, Y. Ikegaya, Inhibition allocates spikes during hippocampal ripples. Nat. Commun. 13, 1280 (2022). doi: 10.1038/s41467-022- 28890-9; pmid: 35277500 L. Muessig, M. Lasek, I. Varsavsky, F. Cacucci, T. J. Wills, Coordinated emergence of hippocampal replay and theta sequences during post-natal development. Curr. Biol. 29, 834–840.e4 (2019). doi: 10.1016/j.cub.2019.01.005; pmid: 30773370 84. 63. G. M. Wittenberg, S. S.-H. Wang, Malleability of spike-timing- dependent plasticity at the CA3-CA1 synapse. J. Neurosci. 26, 6610–6617 (2006). doi: 10.1523/JNEUROSCI.5388-05.2006; pmid: 16775149 85. K. A. Bolding, J. Ferbinteanu, S. E. Fox, R. U. Muller, Place cell firing cannot support navigation without intact septal circuits. Hippocampus 30, 175–191 (2020). doi: 10.1002/hipo.23136; pmid: 31301167 64. R. K. Mishra, S. Kim, S. J. Guzman, P. Jonas, Symmetric spike 86. D. Robbe, G. Buzsáki, Alteration of theta timescale dynamics timing-dependent plasticity at CA3-CA3 synapses optimizes storage and recall in autoassociative networks. Nat. Commun. 7, 11552 (2016). doi: 10.1038/ncomms11552; pmid: 27174042 65. C. Drieu, R. Todorova, M. Zugaro, Nested sequences of hippocampal assemblies during behavior support subsequent sleep replay. Science 362, 675–679 (2018). doi: 10.1126/ science.aat2952; pmid: 30409880 66. U. Farooq, G. Dragoi, Emergence of preconfigured and plastic time-compressed sequences in early postnatal development. Science 363, 168–173 (2019). doi: 10.1126/science.aav0502; pmid: 30630930 of hippocampal place cells by a cannabinoid is associated with memory impairment. J. Neurosci. 29, 12597–12605 (2009). doi: 10.1523/JNEUROSCI.2407-09.2009; pmid: 19812334 87. N. T. M. Robinson et al., Medial entorhinal cortex selectively supports temporal coding by hippocampal neurons. Neuron 94, 677–688.e6 (2017). doi: 10.1016/ j.neuron.2017.04.003; pmid: 28434800 88. N. M. van Strien, N. L. M. Cappaert, M. P. Witter, The anatomy of memory: An interactive overview of the parahippocampal-hippocampal network. Nat. Rev. 38. C. Grienberger, J. C. Magee, Entorhinal cortex directs 59. S. J. Middleton, T. J. McHugh, Silencing CA3 disrupts Liu et al., Science 382, eadi8237 (2023) 20 October 2023 15 of 16 RES EARCH | R E S E A R C H A R T I C L E Neurosci. 10, 272–282 (2009). doi: 10.1038/nrn2614; pmid: 19300446 Neuron 109, 363–376.e6 (2021). doi: 10.1016/ j.neuron.2020.10.034; pmid: 33217328 90. 89. S. Melzer et al., Long-range-projecting GABAergic neurons modulate inhibition in hippocampus and entorhinal cortex. Science 335, 1506–1510 (2012). doi: 10.1126/ science.1217139; pmid: 22442486 J. Basu et al., Gating of hippocampal activity, plasticity, and memory by entorhinal cortex long-range inhibition. Science 351, aaa5694 (2016). doi: 10.1126/science.aaa5694; pmid: 26744409 A. Fernández-Ruiz et al., Long-duration hippocampal sharp wave ripples improve memory. Science 364, 1082–1086 (2019). doi: 10.1126/science.aax0758; pmid: 31197012 92. A. Oliva, A. Fernández-Ruiz, G. Buzsáki, A. Berényi, Spatial 91. coding and physiological properties of hippocampal neurons in the Cornu Ammonis subregions. Hippocampus 26, 1593–1607 (2016). doi: 10.1002/hipo.22659; pmid: 27650887 93. A. Mathis et al., DeepLabCut: Markerless pose estimation of user-defined body parts with deep learning. Nat. Neurosci. 21, 1281–1289 (2018). doi: 10.1038/s41593-018-0209-y; pmid: 30127430 94. M. Stimberg, R. Brette, D. F. Goodman, Brian 2, an intuitive and efficient neural simulator. eLife 8, e47314 (2019). doi: 10.7554/eLife.47314; pmid: 31429824 95. M. Pachitariu, N. A. Steinmetz, S. N. Kadir, M. Carandini, K. D. Harris, Fast and accurate spike sorting of high-channel count probes with KiloSort. Adv. Neural Inf. Process. Syst. 29, 4448–4456 (2016). 96. P. C. Petersen, J. H. Siegle, N. A. Steinmetz, S. Mahallati, G. Buzsáki, CellExplorer: A framework for visualizing and characterizing single neurons. Neuron 109, 3594–3608.e2 (2021). doi: 10.1016/j.neuron.2021.09.002; pmid: 34592168 97. W. Skaggs, B. McNaughton, K. Gothard, An information- theoretic approach to deciphering the hippocampal code. Adv. Neural Inf. Process. Syst. 5, 1030–1037 (1992). 98. S. Leutgeb et al., Independent codes for spatial and episodic memory in hippocampal neuronal ensembles. Science 309, 619–623 (2005). doi: 10.1126/science.1114037; pmid: 16040709 99. C. Dong, A. D. Madar, M. E. J. Sheffield, Distinct place cell dynamics in CA1 and CA3 encode experience in new environments. Nat. Commun. 12, 2977 (2021). doi: 10.1038/ s41467-021-23260-3; pmid: 34016996 100. F. Sharif, B. Tayebi, G. Buzsáki, S. Royer, A. Fernandez-Ruiz, Subcircuits of deep and superficial CA1 place cells support efficient spatial coding across heterogeneous environments. 101. R. Kempter, C. Leibold, G. Buzsáki, K. Diba, R. Schmidt, Quantifying circular-linear associations: Hippocampal phase precession. J. Neurosci. Methods 207, 113–124 (2012). doi: 10.1016/j.jneumeth.2012.03.007; pmid: 22487609 102. K. Diba, G. Buzsáki, Hippocampal network dynamics constrain the time lag between pyramidal cells across modified environments. J. Neurosci. 28, 13448–13456 (2008). doi: 10.1523/JNEUROSCI.3824-08.2008; pmid: 19074018 103. T. J. Davidson, F. Kloosterman, M. A. Wilson, Hippocampal replay of extended experience. Neuron 63, 497–507 (2009). doi: 10.1016/j.neuron.2009.07.027; pmid: 19709631 104. H. S. Kudrimoti, C. A. Barnes, B. L. McNaughton, Reactivation of hippocampal cell assemblies: Effects of behavioral state, experience, and EEG dynamics. J. Neurosci. 19, 4090–4101 (1999). doi: 10.1523/JNEUROSCI.19-10-04090.1999; pmid: 10234037 105. B. Giri, H. Miyawaki, K. Mizuseki, S. Cheng, K. Diba, Hippocampal reactivation extends for several hours following novel experience. J. Neurosci. 39, 866–875 (2019). doi: 10.1523/JNEUROSCI.1950-18.2018; pmid: 30530857 106. V. Lopes-dos-Santos, S. Ribeiro, A. B. L. Tort, Detecting cell assemblies in large neuronal populations. J. Neurosci. Methods 220, 149–166 (2013). doi: 10.1016/j.jneumeth.2013.04.010; pmid: 23639919 107. G. M. van de Ven, S. Trouche, C. G. McNamara, K. Allen, D. Dupret, Hippocampal offline reactivation consolidates recently formed cell assembly patterns during sharp wave- ripples. Neuron 92, 968–974 (2016). doi: 10.1016/ j.neuron.2016.10.020; pmid: 27840002 108. R. Todorova, M. Zugaro, Isolated cortical computations during delta waves support memory consolidation. Science 366, 377–381 (2019). doi: 10.1126/science.aay0616; pmid: 31624215 109. V. J. Oberto et al., Distributed cell assemblies spanning prefrontal cortex and striatum. Curr. Biol. 32, 1–13.e6 (2022). doi: 10.1016/j.cub.2021.10.007; pmid: 34699783 110. A. B. L. Tort et al., Dynamic cross-frequency couplings of local field potential oscillations in rat striatum and hippocampus during performance of a T-maze task. Proc. Natl. Acad. Sci. U.S.A. 105, 20517–20522 (2008). doi: 10.1073/pnas.0810524105; pmid: 19074268 111. N. Benito et al., Spatial modules of coherent activity in pathway-specific LFPs in the hippocampus reflect topology and different modes of presynaptic synchronization. Cereb. Cortex 24, 1738–1752 (2014). doi: 10.1093/cercor/bht022; pmid: 23395845 112. A. J. Bell, T. J. Sejnowski, An information-maximization approach to blind separation and blind deconvolution. Neural Comput. 7, 1129–1159 (1995). doi: 10.1162/ neco.1995.7.6.1129; pmid: 7584893 113. V. Saravanan, G. J. Berman, S. J. Sober, Application of the hierarchical bootstrap to multi-level data in neuroscience. Neuron Behav. Data Anal. Theory 3, 1–25 (2020). pmid: 33644783 114. R. harvey et al., ayalab1/neurocode: Release (v1.0.0), version 1.0.0, Zenodo (2023); https://doi.org/10.5281/ zenodo.7819979. AC KNOWLED GME NTS The authors thank R. Harvey, H. Robison, L. Sjulson, J. Goldberg, I. Ellwood, W. Sun, M. Zugaro, and D. Nitz for useful comments on the manuscript and M. R. Warden for providing access to the confocal microscope. Funding: This work was supported by NIH grant 4R00MH120343, a Sloan Fellowship, a Whitehall Research Grant, and a Klingenstein-Simons Fellowship (A.F.-R.); NIH grant 4R00MH122582 (A.O.); an Eric and Wendy Schmidt AI in Science Postdoctoral Fellowship (a Schmidt Futures program) and a Philippe Foundation grant (R.T.); and a Klarman Fellowship (W.T.). Author contributions: All authors conceived of and designed the experiments and analyses and wrote the manuscript. C.L., W.T., and A.F.-R. performed the experiments. R.T., C.L., W.T., and A.F.-R. analyzed the data. R.T. performed the simulations. First authorship order was determined by coin toss. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data are available in the main text or the supplementary materials. Custom scripts used in this study can be downloaded from Zenodo (114). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi8237 Figs. S1 to S14 MDAR Reproducibility Checklist Submitted 19 May 2023; accepted 21 August 2023 10.1126/science.adi8237 Liu et al., Science 382, eadi8237 (2023) 20 October 2023 16 of 16
10.1126_science.adj0040
RES EARCH ORGANIC CHEMISTRY Multiplicative enhancement of stereoenrichment by a single catalyst for deracemization of alcohols Lu Wen†, Jia Ding†, Lingfei Duan†, Shun Wang, Qing An, Hexiang Wang, Zhiwei Zuo* Stereochemical enrichment of a racemic mixture by deracemization must overcome unfavorable entropic effects as well as the principle of microscopic reversibility; recently, photochemical reaction pathways unveiled by the energetic input of light have led to innovations toward this end, most often by ablation of a stereogenic C(sp3)–H bond. We report a photochemically driven deracemization protocol in which a single chiral catalyst effects two mechanistically different steps, C–C bond cleavage and C–C bond formation, to achieve multiplicative enhancement of stereoinduction, which leads to high levels of stereoselectivity. Ligand-to-metal charge transfer excitation of a titanium catalyst coordinated by a chiral phosphoric acid or bisoxazoline efficiently enriches racemic alcohols that feature adjacent and fully substituted stereogenic centers to enantiomeric ratios up to 99:1. Mechanistic investigations support a pathway of sequential radical-mediated bond scission and bond formation through a common prochiral intermediate and reveal that, although the overall stereoenrichment is high, the selectivity in each individual step is moderate. U biquitous C(sp3)–C(sp3) bonds in organic compounds have recently been exploited as unconventional functional handles for rapid complexity generation through skeletal editing (1–5). This logic presents intriguing opportunities for stereogenic bond construction via asymmetric catalysis. Most commonly, stereogenic C–C bonds are formed by facially selective addition to a prochiral start- ing material (Fig. 1A) (6–8). Critically, the overall enantioselectivity of this process is thus a direct function of the stereodifferentiation of the two prochiral faces in the single stereodefining step [enantiomeric ratio (er) = kR/kS, where kR is the rate of formation of the (R)-enantiomer and kS is the rate of formation of the (S)-enantiomer]. As a result, exceptionally high degrees of stereo- induction are required in this irreversible addi- tion step to achieve synthetically useful results (kR >> kS). An alternative paradigm can be en- visioned transiting through a prochiral inter- mediate to enable a formally reversible C–C bond formation process, which can convert a mixture of racemates into enantiomerically pure com- pounds (Fig. 1B). This approach would enable a broader purview for catalytic deracemization because C–C bonds constitute the fundamental three-dimensional skeleton of complex mole- cules. Through stereoisomerization of the molec- ular core, as opposed to the peripheral C–H bonds, consecutive or fully substituted stereogenic cen- ters could be directly enriched by using this cycle of stereocenter-ablating C–C bond scission and stereocenter-generating C–C bond formation, State Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, University of Chinese Academy of Sciences, Chinese Academy of Sciences, Shanghai 200032, China. *Corresponding author. Email: zuozhw@sioc.ac.cn †These authors contributed equally to this work. which would provide a powerful platform for enantioenrichment via catalytic deracemization. Within this framework, we recently won- dered whether one catalyst could be directly responsible for each of these stereoselective steps by means of different mechanisms, which would thus allow for deracemization. The multi- plicative nature of stereoinduction would lead to amplification of the enantiodifferentation ability of the catalyst, in analogy to the Horeau principle in which multiple asymmetric steps do not necessarily need rigid stereocontrol to obtain highly enantioenriched diastereomers (9–11). However, in this case, because one of the stereodifferentiating steps in the catalytic cycle involves stereoablation, the same multi- plicative enhancement could be achieved in the context of enantioinduction of even a sin- gle stereocenter (er = kRk–S/kSk–R). To achieve synthetically useful levels of enantioinduction [>90% enantiomeric excess (ee)], the single cat- alyst would only be required to achieve roughly 60% ee in each of the stereocenter-ablating and stereocenter-generating steps. Although conceptually simple and synthet- ically intriguing, the development of such pro- cesses has been hindered by the principle of microscopic reversibility (12–18). To address this critical issue, previous studies have opted to use a single chiral catalyst for deracemiza- tion through a two-step process that consists of a nonselective step and a stereodifferentiating step. Notable early studies on redox-driven de- racemizations by Toste et al. (19) and Zhou et al. (20), as well as recent breakthroughs in photo- chemical deracemization such as Bach et al.’s seminal works that used enantioselective trip- let sensitization (18, 21–23) and several other elegant systems with reversible C–H bond formation (24–29), have demonstrated the effectiveness of this approach (30, 31). Very recently, Knowles and Miller and co-workers used an elegantly designed tricatalytic sys- tem that consisted of one chiral base for C–H bond breakage and a chiral thiol catalyst for C–H bond formation, which exploited the syn- ergistic effect of two chiral catalysts for the photocatalytic deracemization of cyclic ureas (32). It is important to note that using one chiral catalyst to effect asymmetric induction in both bond-breaking and bond-making events will result in an overall equilibrium of racemate because of the shared energy surface (33, 34), unless distinct reaction pathways can be re- alized by a single catalyst. Recently, we have applied ligand-to-metal charge transfer (LMCT) excitation in an or- chestrated sequence of in situ coordination, LMCT homolysis, and alkoxy radical–mediated b-scission to the catalytic activation of C–C bonds of free alcohols for fragment couplings (35–38). Driven by the thermodynamic stabil- ity of radicals, both strained and unstrained C–C bonds can be selectively and irreversib- ly cleaved into a carbonyl unit and a carbon- centered radical (39). Notably, high-valent metal ions, including commonly used Lewis acids, can be directly used for LMCT catalysis (40–45), which led us to consider the possibil- ity of photocatalytic deracemization using chiral ligand–coordinated, LMCT-competent Lewis acid catalysts (Fig. 1C). Conceivably, asymmetric LMCT catalysts could leverage the chiral recog- nition of racemic alcohols in the coordination step to preferentially form one diastereomeric metal alkoxide complex, initiating enantiose- lective C–C bond scission, even if the b-scission process may not respond to asymmetric induc- tion. The concurrent generation of a carbonyl fragment and a transient carbon-centered radical would set the stage for single-electron reduction and enantioface-differentiating ad- dition to rebuild the stereogenic C–C bond, which could be facilitated by the lower-valent chiral metal complex generated in the photoexcita- tion event (46–55). In this work, we realize this design plan and describe an LMCT-enabled deracemization platform that uses a single chiral Ti-catalyst to induce decoupled and enantioselective C–C bond cleavage and for- mation, which results in high stereoselectivity through multiplicative enhancement of stereo- induction. By exploiting the photocatalytic properties of a common Lewis acidic Ti(IV) catalyst ligated by chiral phosphoric acid or bisoxazoline, racemic alcohols—including those that feature adjacent and fully substituted ste- reogenic centers—can be efficiently converted into their enantioenriched forms with pro- nounced selectivity. Catalyst optimization with cyclic alcohols We selected the cis isomer of 2-phenylcyclo- pentanol (1) as the model substrate to explore LMCT catalysis for deracemization (Fig. 2). Ex- tensive evaluation of high-valent metal catalysts Wen et al., Science 382, 458–464 (2023) 27 October 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 1. Catalytic deracemization paradigms via enantioselective bond scission and formation. Multiplicative enhancement of two distinct enantioinduction steps enables high levels of stereoselectivities in deracemization. (A) Established enantioselective C–C bond formations. (B) Catalytic deracemization via enantioselective C–C bond scission and formation. (C) Deracemization of free alcohols enabled by asymmetric LMCT catalysis. Bn, benzyl; Boc, t-butyloxy carbonyl; hv, photon energy; L*, chiral ligand; Me, methyl; Ts, p-toluenesulfonyl; Ph, phenyl; Re, R prochiral face. Fig. 2. Reaction optimization. Reactions were performed on a 0.05-mmol scale. Yields, dr, and er (given for major diastereomers) were determined by high-performance liquid chromatography (HPLC) analysis. C6F5, pentafluorophenyl; EtN(iPr)2, diisopropylethylamine; Tf, trifluoromethanesulfonyl; TRIP, 2,4,6-triisopropylphenyl. with chiral ligands under the irradiation of light-emitting diodes (LEDs) (peak wavelength, 395 nm; light intensity, 0.8 W/cm2) at 20°C (see fig. S1 for detailed experimental setup) revealed the combination of TiCl4 and chiral phosphoric acid (CPA) as an effective cata- lyst combination for photocatalytic deracem- ization. This chiral Lewis acid combination has been previously used by Leibfarth to control enantiodifferentiating C–C bond formation for stereoselective cationic polymerization (56). In practice, we found that the combination of 4 mol % TiCl4 and 16 mol % (S)-L1 in the presence of substoichiometric Na2CO3 achieved optimal efficiency and selectivity, generating Wen et al., Science 382, 458–464 (2023) 27 October 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Reaction scope of cycloalkanols. Yields and dr and er values were determined based on isolated products and were the average of three parallel reactions. *4 mol % TiCl4 with 16 mol % L1. †2 mol % TiCl4 with 8 mol % L6. ‡4 mol % TiCl4 with 16 mol % L6. §2 mol % TiCl4 with 8 mol % L5. ¶−40°C. #0°C. See supplementary materials for experimental details. C10F7, heptafluoronaphthyl; C12F9, nonafluorobiphenyl; Et, ethyl. cis (1S,2S)-1 with high efficiency and notable enantioselectivity (93% yield, 99:1 er, entry 1). Under these mild and redox-neutral condi- tions, only a small amount of trans isomer was obtained along with the desired cis isomer [8.8:1 diastereomeric ratio (dr)], without the observation of side products such as aldehyde from photocatalytic ring-opening processes, ketone from 2e oxidations, or alkene from de- hydration (see fig. S2). The steric environment of the CPA has a notable influence on the stereo- chemical outcome. Replacing the sterically hin- dered tri-iso-propylphenyl rings of L1 with trimethylphenyl (L2) resulted in markedly diminished stereoselectivity (entry 2), where- as phosphoric acids bearing electron-deficient arene substituents (L3, L4, L5, or L6) re- sulted in rather low selectivities (entries 3 to 6). Use of the more acidic but less coordinat- ing N-triflyl phosphoramide L7 in place of L1 resulted in markedly lower stereoselectivity (3.2:1 dr, 21:79 er) despite similar steric envi- ronments (entry 7). Organic base diisopropyl- ethylamine was also found to be effective for the desired deracemization (entry 8). Chang- ing the loading of CPA relative to Ti from 4:1 to 2:1 had little effect on the efficiency and stereoselectivity (see table S3), and the 4:1 ratio was adopted for best reproducibility on small-scale parallel reactions. A linear corre- lation between the enantioenrichment ob- served in the reaction and the ee value of CPA was identified (see fig. S4), and the absence Wen et al., Science 382, 458–464 (2023) 27 October 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 4. Deracemization of acyclic alcohols. Yields and dr and er values were determined on the basis of isolated products and were the average of three parallel reactions. *8 mol % TiCl4 with 20 mol % L8. †12 mol % TiCl4 with 30 mol % L8. ‡Dichloromethane (DCM) and benzotrifluoride (PhCF3) (1:1). §PhCF3 instead of DCM. of a nonlinear effect in this deracemization indicated that a single CPA is presumably in- volved at the Ti center (57). Use of (R)-L1 re- sulted in identical values of stereoselectivity with opposite configuration, supporting the catalyst-controlled stereoselectivity (entry 9). Furthermore, deracemization of the second- ary alcohol was not observed when swapping TiCl4 with CeCl3 or FeCl3 (entries 10 and 11), and instead, a trace amount of 5-phenylpentanal was observed, indicating that Ce and Fe LMCT catalysts are capable of C–C bond scission but not effective at promoting C–C bond–forming addition. Moreover, control experiments re- vealed that L1, TiCl4, and an LED light are essential for the deracemization (entries 12 to 14). As demonstrated in Fig. 3, this Ti-LMCT– enabled photocatalytic deracemization can be carried out on structurally diverse cyclo- alkanols of varying ring sizes, indicating that ring strain is not a decisive factor. In cyclo- pentanol and cyclohexanol substrates, we found that the deracemization is somewhat insen- sitive to the electronic nature of the b-aromatic rings. Excellent enantioselectivities could be obtained for a series of diverse substituents on the para, ortho, and meta positions, with trivial differences in the diastereoselectivities. Even in the presence of large aromatic rings such as naphthalene (11), benzothiophene (12), and dibenzofuran (13), the thermodynamical- ly less-stable cis isomers were preferentially formed for cyclopentanols. Regarding the six- membered ring system, enantioenriched trans isomers were selectively generated, presumably by virtue of a chair-like cyclic transition state for C–C bond formation. We found that (R)- L6 was most effective for the deracemization of trans-2-substituted cyclohexanols. Notably, enantioenriched 3-hydroxy piperidine (23), an essential building block for the syntheses of important marketed pharmaceuticals such as zamifenacin, piperidolate, and benidipine, can also be selectively obtained with this straight- forward protocol (58). Critically, enantioen- riched synthesis of a-tertiary alcohols with fully substituted stereogenic carbons could be facilely achieved by this deracemization plat- form. With the catalytic combination of TiCl4 and (S)-L5, a variety of tertiary 3-hydroxy pi- peridines could be obtained with high levels of enantioenrichment, regardless of the elec- tronic nature of the a-aromatic rings. Tertiary alcohols with a-methyl substituents can also be accommodated, which renders structurally diversified piperidine scaffolds with excellent stereoselectivity. Further studies revealed that larger alkyl groups at the a-position, including ethyl and isobutyl groups, resulted in somewhat lower er (see table S10). The stereochemistry of tertiary alcohols 24 and 34 has been unam- biguously assigned by single-crystal x-ray dif- fraction. Even b-substituted cyclobutanols were compatible with this deracemization paradigm and generated enantioenriched cyclobutanols 35 to 38 with good enantioselectivities through enantioselective C–C bond scission and reforma- tion of the highly strained four-membered ring. The generation of aldehyde byproducts through a photocatalytic ring-opening process was ob- served in those strained substrates, which re- sulted in declined yields of deracemization. Secondary and tertiary alcohols embedded in seven-membered carbocycle or azepane scaf- folds were also compatible, albeit with mod- erate selectivities. We conducted a scaled-up deracemization of cyclohexanol 14 at 1 mmol scale by using a simple and easily assembled photoflow system (see section 5.1 in the sup- plementary materials), which rendered (−)-14 with similar levels of selectivities as the stan- dard conditions (91% yield, 7.2:1 dr, 95:5 er). Extension to acyclic alcohols This protocol is not limited to cyclic hydroxyl- ated scaffolds because the deracemization of acyclic alcohols has been validated on 1,2-diaryl aminoalcohols, a privileged scaffold for chiral ligand synthesis (Fig. 4). Under similar cata- lytic conditions as for the cyclic substrates, with chiral bisoxazoline L8 as the ligand (see table S11 for the ligand evaluation), enantioenriched (+)-41 can be obtained with excellent yield and stereoselectivity (20:1 dr, 98:2 er). A variety of substitution patterns on the a- and b-aromatic rings proved compatible, and high yields and excellent stereoselectivities have been obtained with no sensitivity to the electronic property of the aromatic rings. This deracemization has provided a practical approach to enantio- enriched aminoalcohols with electronically differentiated aryl rings on two different car- bon termini (59, 60). Moreover, we conducted 1-mmol-scale reactions with 41 and 43, which delivered enantiopure alcohols with identi- cal selectivities to the standard condition at prolonged reaction time (10 hours). Compound (+)-43 can be elaborated via deprotection and double condensation to generate a differen- tially arylated bisoxazoline ligand, the para- chloro substituted version of L8 (see section 5.2 in the supplementary materials). Mechanistic investigations We carried out mechanistic investigations to elucidate the reaction pathway. Using C–H- deuterated cylopentanol 50 and aminoalcohol 51 delivered nearly identical stereochemical outcomes as the unlabeled analogs without eroding the deuteration ratio, which precludes hydrogen atom transfer (HAT) or stepwise oxi- dation reduction as possible deracemization pathways (Fig. 5A). With the same ligand, en- antioconvergent transformations that used the trans isomer of 1 or the opposite enantiomer (1R,2R)-1 resulted in the same (1S,2S)-1 product with identical stereoselectivity, supporting a common, prochiral intermediate generated by the C–C bond scission (Fig. 5B). This conclusion can also be drawn from the same set of exper- iments with aminoalcohol 41 (fig. S8). The LMCT-homolysis behavior of Ti(IV) com- plexes has been validated by steady-state photol- ysis experiments by using an in situ–generated putative complex [Ti(IV)L(OR)] (Fig. 5C). We used a mononuclear [TiIV(L8)Cl4] complex, char- acterized by x-ray diffraction, in the ligand Wen et al., Science 382, 458–464 (2023) 27 October 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E Fig. 5. Preliminary mechanistic investigations. (A) Isotope labeling experiments. (B) Enantioconvergent transformations. (C) Steady-state photolysis experiments. a.u., arbitrary unit; GC-MS, gas chromatography–mass spectrometry. (D) Spin-trapping EPR experiments. Exp, experimental; Sim, simulation. (E) Crossover experiment. (F) Time-course study. (G) Decoupled radial-mediated bond formation and photocatalytic bond scission. THF, tetrahydrofluoran. Wen et al., Science 382, 458–464 (2023) 27 October 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E exchange with an equimolar amount of 46 for the in situ generation of [TiIV(L8)(OR)]. The 1H nuclear magnetic resonance (NMR) spec- trum displayed characteristic signals that can be assigned to the bisoxazoline L8 and alk- oxide ligands in a 1:1 ratio, and diffusion or- dered spectroscopy (DOSY) NMR experiments confirmed that the obtained species are mono- meric complex and not higher aggregates (fig. S10). In the same vein, the putative complex generated from TiCl4 with an equimolar amount of L1 and 14 was prepared (fig. S12). Both complexes exhibited substantial absorptions in the proximity of 395 nm. Under the irra- diation of a 395-nm laser, with the gradual decay of the absorption band ranging from 300 to 400 nm, a weakly absorbing band in the 500- to 800-nm region started to emerge as the originally light yellow solution turned blue, which is indicative of the generation of Ti(III) species (43). Upon exposure to air, the blue solution was swiftly bleached to the yellow Ti(IV) complex with the ultraviolet to visible spectrum fully recovered. Aldehyde 52, formed by radical-mediated bond scission, was detected by gas chromatography–mass spectrometry and 1H NMR, supporting the generation of alkoxy radicals. Regarding the cycloalkanols, the proposed alkyl radical gen- eration by rapid alkoxy radical b-scission was monitored by operando electron paramagnetic resonance experiments with 5,5-dimethyl-1- pyrroline N-oxide (DMPO) and a-phenyl-N-tert- butylnitrone (PBN) as spin-trapping reagents. As depicted in Fig. 5D, the characteristic sig- nals of alkyl radical adduct with DMPO [hyper- fine splitting constant AN = 14.4 G, AH = 20.9 G, spectroscopic splitting factor (g) = 2.0061] or PBN (AN = 14.6 G, AH = 2.4 G, g = 2.0062) can be clearly distinguished, whereas alkoxy radical ad- duct was not detected, indicating a fast b-scission process that favors the C–C bond cleavage or a concerted oxidation with C–C bond scission (61, 62). Moreover, we further confirmed the alkyl radical generated in the b-scission by the isolation and characterization of the 2,2,6,6- tetramethylpiperidin-1-oxyl (TEMPO) adduct. We rationalized that within the same mech- anistic framework, the aromatic aldehyde gen- erated from the b-scission process of acyclic aminoalcohols could be exchanged with a more reactive exogenous aldehyde for the C–C bond formation event to generate another aminoalcohol product. Indeed, under the same catalytic conditions for deracemization, cross- over experiments with racemic 46 and para- fluorobenzaldehyde 53 resulted in a newly formed aminoalcohol 42 that contained a parafluorophenyl group (Fig. 5E). The newly generated aminoalcohol 42 underwent de- racemization to reach the same level of en- antioenrichment as the standard condition (99:1 er). The yield of 42 can be further en- hanced at the expense of deracemization product 46 with increased concentration of aldehyde 53. The foregoing results collec- tively and unambiguously support the inter- mediacy of aldehydes and carbon-centered radicals generated by b-scission as the shared intermediate in this deracemization paradigm. The high efficiency of the desired cyclobutane ring formation in the deracemization of cyclo- butanols is notable because direct intramolec- ular radical addition to the carbonyl to form a four-membered ring is energetically unfa- vorable, considering the high exothermicity and low activation barrier of cyclobutoxy cleav- age (39, 63). There are two seemingly reason- able pathways for this cyclization that account for the ability of Ti(III) to mediate such a rad- ical bond formation: (i) reductive capture of the formed alkyl radical by photogenerated Ti(III) preceding a stereodifferentiating 1,2-addition with the carbonyl group via a six-membered cyclic transition state; or (ii) reversible reduction of the carbonyl to a transient Ti-ketyl radical followed by radical-radical coupling. Although the relative 1,2-stereochemistry observed for cyclobutanol (trans), cyclopentanol (cis), and cyclohexanol (trans) substrates is reminiscent of pseudoequitorial positioning of bulky groups in cyclic transition states (fig. S18), the precise nature of this ring closure warrants further investigation. Insight into the level of asymmetric induc- tion incurred in the C–C bond–forming events can be gleaned by tracking the time course of the deracemization of the syn-46, which re- sults in enantioenriched anti-46 (Fig. 5F). In the first 20 min, the bond cleavage of syn-46 and bond formation step to generate anti-46 results in a steady er value of 75:25. As the concentration of newly generated anti-46 ap- proaches that of syn-46, the er value of anti- 46 gradually increases as the deracemization of the formed major diastereomer takes place, eventually reaching an equilibrium value of 97:3 er. The steady er value of anti-46 during the initial 20 min indicates that the newly formed anti-46 is not yet substantially in- volved in the bond cleavage cycle, and the enantiofacial-determining ratio of the bond formation step can be estimated as 75:25, which results in kR/kS ratio of 3:1. This approximate value is supported by a 77:23 er (kR/kS ratio of 3.3:1) obtained for the inefficient reductive cou- pling of aldehyde 52 and imine 54 under dark conditions with L8-ligated TiCl3, despite pre- dominant pinacol coupling (Fig. 5G) (51, 53, 64). Conversely, we could directly probe the bond scission of aminoalcohol 46 by the introduc- tion of a large amount of radical trapping re- agent 55, which would lead to kinetic resolution based on enantioselective C–C bond cleavage. The asymmetric induction of the C–C bond scission step can be established as a k−S/k−R ratio of 8.1:1, with the (S)-enantiomer preferen- tially consumed. The predicated er value based on the multiplicative effect (er = kRk–S/kSk–R) of these two isolated processes is 96:4, in ex- cellent agreement with the experimental value in the deracemization reaction (97:3) (see fig. S17 for a discussion on cyclic alcohol 1). The LMCT catalytic paradigm enables the use of one chiral Ti catalyst to induce two enantio- selective events in a single catalytic cycle. The multiplicative enhancement of stereoinduction, validated in consecutive stereocenter-ablating C–C bond scission and stereocenter-generating C–C bond formation, enables high levels of stereoselectivities with two moderate enantio- inductions. This mechanistic paradigm holds promise for the development of a wide array of asymmetric catalytic reactions through the statistical upgrading of stereoenrichment. REFERENCES AND NOTES 1. G. Dong, N. Cramer, C−C Bond Activation, vol. 346 of Topics in Current Chemistry, G. Dong, Ed. (Springer, 2014). 2. L. Souillart, N. Cramer, Chem. Rev. 115, 9410–9464 (2015). 3. B. Wang, M. A. Perea, R. Sarpong, Angew. Chem. Int. Ed. 59, 18898–18919 (2020). 4. X.-Y. Yu, J.-R. Chen, W.-J. Xiao, Chem. Rev. 121, 506–561 (2021). J. Jurczyk et al., Nat. Synth. 1, 352–364 (2022). 5. 6. P. J. Walsh, M. C. Kozlowski, Fundamentals of Asymmetric Catalysis (University Science Books, 2009). 7. M. Shibasaki, M. Kanai, Chem. Rev. 108, 2853–2873 (2008). 8. D. J. Ramón, M. Yus, Chem. Rev. 106, 2126–2208 (2006). J. P. Vigneron, M. Dhaenens, A. Horeau, Tetrahedron 29, 9. 1055–1059 (1973). 10. A. M. Harned, Tetrahedron 74, 3797–3841 (2018). 11. M. Burns et al., Nature 513, 183–188 (2014). 12. A. E. Wendlandt, Science 366, 304–305 (2019). 13. Q. Shi, J. Ye, Angew. Chem. Int. Ed. 59, 4998–5001 (2020). 14. M. Huang, T. Pan, X. Jiang, S. Luo, J. Am. Chem. Soc. 145, 10917–10929 (2023). 15. J. Großkopf, T. Bach, Angew. Chem. Int. Ed.e202308241 (2023). 16. A. Tröster, A. Bauer, C. Jandl, T. Bach, Angew. Chem. Int. Ed. 58, 3538–3541 (2019). 17. L. Wimberger, T. Kratz, T. Bach, Synthesis 51, 4417–4416 (2019). 18. X. Li et al., Angew. Chem. Int. Ed. 59, 21640–21647 (2020). 19. A. D. Lackner, A. V. Samant, F. D. Toste, J. Am. Chem. Soc. 135, 14090–14093 (2013). 20. Y. Ji, L. Shi, M.-W. Chen, G.-S. Feng, Y.-G. Zhou, J. Am. Chem. Soc. 137, 10496–10499 (2015). 21. A. Hölzl-Hobmeier et al., Nature 564, 240–243 (2018). 22. M. Plaza, J. Großkopf, S. Breitenlechner, C. Bannwarth, T. Bach, J. Am. Chem. Soc. 143, 11209–11217 (2021). 23. T. Kratz et al., J. Am. Chem. Soc. 144, 10133–10138 (2022). 24. J. Großkopf et al., J. Am. Chem. Soc. 143, 21241–21245 (2021). 25. R. J. Kutta et al., J. Am. Chem. Soc. 145, 2354–2363 (2023). 26. C. Zhang et al., J. Am. Chem. Soc. 143, 13393–13400 (2021). 27. M. Huang, L. Zhang, T. Pan, S. Luo, Science 375, 869–874 (2022). 28. Z. Gu et al., Angew. Chem. Int. Ed. 61, e202211241 (2022). 29. Q. Chen et al., Chem. Sci. 14, 1715–1723 (2023). 30. M. Wan, S. Sun, Y. Li, L. Liu, Angew. Chem. Int. Ed. 56, 5116–5120 (2017). 31. Z. Zhang, X. Hu, Angew. Chem. Int. Ed. 60, 22833–22838 (2021). 32. N. Y. Shin, J. M. Ryss, X. Zhang, S. J. Miller, R. R. Knowles, Science 366, 364–369 (2019). 33. D. G. Blackmond, Angew. Chem. Int. Ed. 48, 2648–2654 (2009). 34. W. Kroutil, K. Faber, Tetrahedron Asymmetry 9, 2901–2913 (1998). 35. J.-J. Guo et al., Angew. Chem. Int. Ed. 55, 15319–15322 (2016). 36. A. Hu et al., J. Am. Chem. Soc. 140, 13580–13585 (2018). 37. K. Zhang, L. Chang, Q. An, X. Wang, Z. Zuo, J. Am. Chem. Soc. 141, 10556–10564 (2019). Wen et al., Science 382, 458–464 (2023) 27 October 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E 38. Y. Chen, X. Wang, X. He, Q. An, Z. Zuo, J. Am. Chem. Soc. 143, 54. T. McCallum, X. Wu, S. Lin, J. Org. Chem. 84, 14369–14380 4896–4902 (2021). 39. L. Chang, Q. An, L. Duan, K. Feng, Z. Zuo, Chem. Rev. 122, 2429–2486 (2022). (2019). 55. X. Wu, Y. Chang, S. Lin, Chem 8, 1805–1821 (2022). 56. A. J. Teator, F. A. Leibfarth, Science 363, 1439–1443 40. S. M. Treacy, T. Rovis, J. Am. Chem. Soc. 143, 2729–2735 (2019). (2021). 41. Z. Zhang et al., Angew. Chem. Int. Ed. 59, 9355–9359 (2020). 42. M. Yamane, Y. Kanzaki, H. Mitsunuma, M. Kanai, Org. Lett. 24, 1486–1490 (2022). 43. A. Fermi, A. Gualandi, G. Bergamini, P. G. Cozzi, Eur. J. Org. Chem. 2020, 6955–6965 (2020). 44. F. Juliá, ChemCatChem 14, e202200916 (2022). 45. C. Wegeberg, O. S. Wenger, JACS Au 1, 1860–1876 (2021). 46. A. Gansäuer, L. Shi, M. Otte, J. Am. Chem. Soc. 132, 11858–11859 (2010). 47. J. Streuff, M. Feurer, P. Bichovski, G. Frey, U. Gellrich, Angew. Chem. Int. Ed. 51, 8661–8664 (2012). 48. W. Hao, J. H. Harenberg, X. Wu, S. N. MacMillan, S. Lin, J. Am. Chem. Soc. 140, 3514–3517 (2018). 49. A. Gualandi et al., ACS Catal. 10, 3857–3863 (2020). 50. F. Li et al., Angew. Chem. Int. Ed. 60, 1561–1566 (2021). 51. F. Calogero et al., Chem. Sci. 13, 5973–5981 (2022). 52. X. Peng et al., J. Org. Chem. 88, 6333–6346 (2023). 53. B. Rossi, S. Prosperini, N. Pastori, A. Clerici, C. Punta, Molecules 17, 14700–14732 (2012). 57. T. Satyanarayana, S. Abraham, H. B. Kagan, Angew. Chem. Int. Ed. 48, 456–494 (2009). 58. C. J. Wenthur et al., J. Med. Chem. 56, 5208–5212 (2013). 59. A. Z. Gonzalez et al., J. Med. Chem. 57, 2472–2488 (2014). 60. S. K. Gediya, G. J. Clarkson, M. Wills, J. Org. Chem. 85, 11309–11330 (2020). 61. Q. An et al., J. Am. Chem. Soc. 145, 359–376 (2023). 62. S. Kim, P.-P. Chen, K. N. Houk, R. R. Knowles, J. Am. Chem. Soc. 144, 15488–15496 (2022). 63. S. Wilsey, P. Dowd, K. N. Houk, J. Org. Chem. 64, 8801–8811 (1999). 64. A. Bensari, J.-L. Renaud, O. Riant, Org. Lett. 3, 3863–3865 (2001). ACKN OWLED GMEN TS Funding: We thank the National Natural Science Foundation of China (grant nos. 22125111 and 21971163), National Key R&D Program of China (grant no. 2021YFA1500100), the Shanghai Pilot Program for Basic Research – Chinese Academy of Science, Shanghai Branch, and SIOC for financial support. We also thank J. Lipshultz for thoughtful discussions and J. Wu for assistance with the NMR experiments. Author contributions: L.W., J.D., and L.D. contributed equally to this work. Z.Z. conceived and directed the research; Z.Z., L.W., J.D., L.D., and S.W. designed the experiments. L.W., J.D., L.D., S.W., Q.A., and H.W. performed and analyzed the reactions; Z.Z., L.W., and S.W. prepared the manuscript, which was approved by all authors. Competing interests: The authors declare no conflicts of interest. Data and materials availability: Data are available in the manuscript and supplementary materials. The supplementary crystallographic data for this paper are available free of charge from the Cambridge Crystallographic Data Centre (CCDC) under accession numbers 2265923, 2265930, 2266206, and 2287759. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adj0040 Materials and Methods Supplementary Text Figs. S1 to S20 Tables S1 to S42 References (65–69) Submitted 1 June 2023; accepted 4 September 2023 10.1126/science.adj0040 Wen et al., Science 382, 458–464 (2023) 27 October 2023 7 of 7
10.1126_science.adi8474
RES EARCH MACHINE LEARNING Backpropagation-free training of deep physical neural networks Ali Momeni1, Babak Rahmani2, Matthieu Malléjac1, Philipp del Hougne3, Romain Fleury1* Recent successes in deep learning for vision and natural language processing are attributed to larger models but come with energy consumption and scalability issues. Current training of digital deep-learning models primarily relies on backpropagation that is unsuitable for physical implementation. In this work, we propose a simple deep neural network architecture augmented by a physical local learning (PhyLL) algorithm, which enables supervised and unsupervised training of deep physical neural networks without detailed knowledge of the nonlinear physical layer’s properties. We trained diverse wave-based physical neural networks in vowel and image classification experiments, showcasing the universality of our approach. Our method shows advantages over other hardware-aware training schemes by improving training speed, enhancing robustness, and reducing power consumption by eliminating the need for system modeling and thus decreasing digital computation. D eep learning has emerged as a break- through technology with outstanding success (1, 2) that primarily operates on traditional von Neumann computing hardware. This technology is currently facing high energy consumption, such as the 1.3–gigawatt-hour (GWh) electricity usage of GPT-3 (3), and low computing speed (4). Be- cause of these challenges, researchers are exploring alternative physical platforms for artificial neural networks (ANNs), including optics (5–9), spintronics, (10, 11), nanoelectronic devices (12–15), photonic hardware (5), and acoustic systems (16, 17). Two primary methods currently dominate neural network hardware design. The first in- volves designing hardware to implement trained mathematical transformations through strict operation-by-operation mathematical iso- morphism, primarily targeting the inference phase of deep learning (18–21). The second category, deep physical neural networks (PNNs), focuses on training the hardware’s physical transformations directly to perform the de- sired computations. PNNs hold the promise of more scalable, energy-efficient, and faster neural network hardware by exploiting phys- ical transformations and eliminating the conventional software-hardware separation (22, 23). So far, the training of PNNs has predom- inantly relied on backpropagation (BP) (24). Yet, there are several reasons why BP is not a suitable choice for PNNs, one of which is the complexity and lack of scalability in the phys- ical implementations of BP operations in the 1Laboratory of Wave Engineering, Department of Electrical Engineering, EPFL, Lausanne CH-1015, Switzerland. 2Microsoft Research, Cambridge CB4 0AB, UK. 3University of Rennes, CNRS, IETR - UMR 6164, F-35000 Rennes, France. *Corresponding author. Email: romain.fleury@epfl.ch hardware (25–28). Commonly, PNN proposals use in silico training, performing BP calcu- lations on an external computer with a digital twin of the physical system. However, this meth- od may result in potential simulation-reality gaps as a result of inaccurate representation of the physical system (6–8, 10, 13, 14, 20, 29, 30). Moreover, physics-aware training methods based on BP (PA-BP) (22) offer improvements over traditional in silico methods but still neces- sitate a differentiable digital model for the backward pass. Additionally, PA-BP–trained PNNs may face challenges when subjected to strong perturbations, potentially rendering fine- tuned models unusable and necessitating re- training from scratch. Another important drawback of BP is its re- liance on having complete knowledge of the computation graph carried out during the forward pass to accurately compute deriva- tives (23, 31–34). When a black box is inserted in the forward pass, BP becomes infeasible. Therefore, alternative training methods for PNNs have proved advantageous. For exam- ple, an approach explored for training physical networks is the augmented direct feedback alignment (DFA) method (23), which aims to avoid the need for a differentiable digital mod- el. However, this method is only compatible with certain physical networks, where it is possible to separate the nonlinear and linear layers. Local learning has been extensively studied for training digital neural networks, from early work on Hebbian contrastive learning in Hopfield models (35) to recent biologically plausible frameworks (31, 34, 36, 37), block- wise BP (38, 39), and contrastive representa- tion learning (40, 41). Inspired by this concept and to address the limitations of BP-based PNN training, we proposed a simple and physics- compatible PNN architecture augmented by a physical local learning (PhyLL) algorithm. The proposed method enables supervised and unsupervised contrastive learning training of arbitrary PNNs locally without the need to know the nonlinear physical layers and train a digital twin model. In this BP-free method, the standard backward pass, typically performed by a digital computer, is replaced with an ad- ditional single forward pass through a physical system. This substitution can improve training speed, power consumption, and memory usage during the training phase of wave-based PNNs by eliminating the extra overhead incurred because of the digital twin modeling phase present in other hardware-aware frameworks. We showed the robustness and adaptability of the proposed method, even in systems ex- posed to unpredictable external perturbations. To showcase the universality of our approach, we performed experimental vowel and image classification using three wave-based systems that differ in terms of the underlying wave phenomenon and the type of nonlinearity in- volved (a detailed description of each system can be found in the supplementary text, sec- tions 2.3 to 2.5). PhyLL Figure 1A shows a simple and physics-compatible deep PNN including N nonlinear physical data transformers augmented by trainable linear multiplications. Each nonlinear phys- ical data transformer performs a nonlinear mapping between the input and output fol- lowed by an augmented trainable linear multi- plication to classify distinct classes through a local training algorithm. The output of each layer is then passed to the next layer. The sub- sequent layer then carries out the same pro- cess hierarchically on the output of the previous layer. The proposed architecture shares some similarities with conventional deep reservoir computing systems (42); see supplementary text, section 2.9, for further details on their differences. h The training algorithm is inspired by the recently proposed forward-forward algorithm (31) and local training proposals (38–41) in digital neural networks, which has been ex- tended and adapted to the supervised and un- supervised model-free physical learning of PNNs. Each nonlinear physical system performs a non- linear transformation on input data (Fig. 1), i which can be expressed as h lð Þ ¼ f lð Þ , N where x lð Þ, W lð Þ N correspond to the physical inputs (e.g., optical intensity, electric voltage, and vibration), the physical intercon- nections (e.g., optical, electrical, or mechan- ical coupling) in the physical system, and the physical nonlinearity (e.g., nonlinear opti- cal, magnetic, or mechanical effects) in layer l, respectively. Here, W lð Þ N denote the mixing oper- ation and nonlinear kernel of the l−th physical p , and f lð Þ p and f lð Þ p x lð Þ W lð Þ Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E t systems, respectively (supplementary text, sec- tions 2.6 and 2.7). The output of layer l can be expressed as the multiplication of h lð Þ by the augmented trainable weight matrix W lð Þ —i.e., t y lð Þ ¼ W lð Þ h lð Þ. Such trainable matrix multi- plications can be performed either digitally or through physical systems, for instance using Mach-Zehnder interferometer (MZI)–integrated photonics (43) or spatial light modulators (SLMs) in optics (21, 44). The goal is to train W lð Þ locally without the need to know the non- t linear physical layer. Instead of a forward and backward pass, we use two physical forward passes: a positive and a negative forward pass through the physical system, each running on different physical inputs. The positive physical W lð Þ pass, y lð Þ i , uses positive pos ¼ W lð Þ p x lð Þ pos h f lð Þ N t t neg f lð Þ N p x lð Þ h W lð Þ i , uses negative inputs inputs that include the input dataset and the correct labels, and the negative physical pass, neg ¼ W lð Þ y lð Þ that include the input dataset and the incor- rect labels (Fig. 1A). In each layer, we calculate the so-called goodness function, defined as the cosine similarity between the positive and negative activities. Eventually, for each layer l, W lð Þ is trained by minimizing the following t loss function L lð Þ ¼ log 1 þ exp q cossim ypos; yneg f ð ð ½ Þ (cid:2) g Þ ð1Þ In supervised learning, the goodness func- tion is defined as the cosine similarity between the activities of the layer and a random vector drawn from normal distribution, both for the Deep physical neural networks positive and negative physical passes. In this case, the loss function reads h n(cid:2) L lð Þ ¼ log 1 þ exp (cid:3)q cossim ypos; x lð Þ h h cossim yneg; x lð Þ o(cid:3)i i i (cid:3) ð2Þ In the equations above, cossim is the cosine similarity defined as the cosine of the angle between the two arguments, q is a scale factor, and xl is the random vector for the layer l. The original forward-forward algorithm uses only the difference of the positive and negative squared activities, hence necessitating layer normalization to be applied to the data before proceeding to subsequent layers (31). Converse- ly, our algorithm avoids incorporating layer A Positive data Correct labels Input Negative data Incorrect labels Input B Acoustics Input Sources Nonlinear membranes Parameters Parameters Nonlinear physical data transformer Trainable linear multiplication g n i t a d p U Output 1 Goodness Local loss Physical data transformers C Microwave Output Metasurface g n i t a d p U Output N Goodness Local loss D Optics Laser Antennas Input (Metasurface config.) Output (Transfer func. Intensity) SLM Input Lenses MMOC CCD Rigid scatterers Frequency Output Fig. 1. Deep PNNs. (A) A simple and physics-compatible deep neural network that uses a sequence of nonlinear physical data transformers augmented by trainable matrix multiplications, trained by the supervised PhyLL technique (refer to supplementary text, section 2.1.1, for additional explanations). At each layer, the nonlinear physical data transformer conducts nonlinear mapping between input and output spaces to separate positive and negative data by maximizing the cosine similarity of the positive data to a random vector x and minimizing the cosine similarity of the negative data to the same vector. We considered three physical systems that vary in terms of the underlying wave phenomenon and the type of nonlinearity. (B) In acoustics, input data are encoded into the intensity of sound waves at different frequencies injected on the left side of the cavity. Sound waves propagate through a chaotic cavity that comprises multiple rigid cylindrical diffusers and nonlinear membranes. The transformed waveforms are received by multiple microphones. (C) In the chaotic microwave cavity, input data are encoded into the programmable metasurface configuration inside the metallic disordered cavity. The outputs are obtained from the waves’ spectra (transfer function). (D) In the optical setup, input data are encoded onto the SLM, and after passing through a multimodal optical cavity (MMOC), the resulting optical intensity is measured on the charge-coupled device (CCD) camera [numerical experiment based on experimentally acquired data from Rahmani et al. (56)]. Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E normalization into the architecture because its hardware-based implementation is a deep challenge. During the inference phase, we in- put a particular label into the PNNs and ac- cumulate the goodness values for all layers. This process is repeated for each label sepa- rately. The label with the highest accumulated goodness value is then selected as the output (see supplementary text, section 2.1, for more details). In unsupervised contrastive learning, a single linear layer maps representations from pretrained hidden layers to labels (supplemen- tary text, section 2.1.1). For details on data gen- eration, refer to supplementary text, section 2.2. Diverse PNNs for vowel and image classification Figure 1 presents three deep PNN classifiers for various standard datasets, including vowel, A B Input layer Physical layer Topology of PNN C 100 y c a r u c c a t s e T 90 80 70 60 50 40 30 20 10 In silico Ideal BP PhyLL 0 50 100 Epoch 150 200 NL membrane Rigid scatterers Sources D 100 y c a r u c c A 90 80 70 60 50 40 30 20 0 50 100 Epoch 150 98.88 % 97.31 % E s e u a v l e u r T Train Test 200 ae ah aw uw er iy ih ae ah aw uw Predicted values er F s e u a v l e u r T ae ah aw uw er iy ih iy ih ae ah 1.0 0.8 0.6 0.4 0.2 0.0 iy ih aw uw Predicted values er Fig. 2. Acoustic-PNN. (A) The topology of the acoustic-PNN consists of a two-layer PNN with skip connections. Each layer comprises an acoustic-PNN augmented by trainable matrix multiplication. (B) Photograph of the experimental setup. NL, nonlinear. (C) Comparison of test accuracy versus training epoch with in silico, ideal BP, and PhyLL algorithm for the vowel recognition task. (D) The train and test classification accuracy versus training epoch for the vowel recognition task. (E and F) The confusion matrix for the PNN on the train (E) and test (F) sets. A C 100 y c a r u c c A 90 80 70 60 50 40 30 Input layer Physical layer Topology of PNN 98,74 % 97.31 % D s e u a v l e u r T Train Test aeae ahah awaw uwuw erer iyiy ihih 0 20 40 60 Epoch 80 100 aeae ahah B Programmable Metasurface Antennas Disorder E s e u a v l e u r T ae ah aw uw er iy ih iyiy ihih ae ah 1.0 0.8 0.6 0.4 0.2 0.0 iy ih aw uw Predicted values er awaw uwuw Predicted values erer Fig. 3. Microwave-PNN. (A) The topology of the microwave-PNN consists of a three-layer PNN with skip connections. Each layer comprises a microwave-PNN augmented by trainable matrix multiplication. (B) Photograph of the experimental setup. (C) Train and test classification accuracy versus training epoch for the vowel recognition task. (D and E) Confusion matrix for the PNN on the train (D) and test (E) sets. Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 3 of 7 RES EARCH | R E S E A R C H A R T I C L E digit, fashion Mnist, and CIFAR10, based on three distinct physical systems, each featur- ing a distinctive source of nonlinearity (ma- terials and methods and supplementary text, sections 2.3 to 2.5). Although there have been proposals that explore wave-based analog com- puting for linear operations, such as multi- plication and convolution (43, 45–53), it is important to note that PNNs require non- linearity to effectively handle regression and classification tasks. We evaluated the per- formance of PhyLL in these media (refer to tables S1 and S2 in the supplementary mate- rials) against in silico and BP methods under both supervised and unsupervised contras- tive training schemes using an end-to-end surrogate forward model of the systems for benchmarking purposes. Acoustic chaotic cavity with nonlinear scatterers In acoustics, an air-filled multimode cavity composed of multiple nonlinear meta-scatterers randomly placed on the cavity top wall and multiple rigid scatterers inside the cavity was used (materials and methods and supplemen- tary text, section 2.3). The nonlinear meta- scatterers were designed to provide a nonlinear relation between pressure and particle velocity with controllable power law. The positive and negative data were encoded onto the ampli- tude of each frequency component composing the excitation waveforms, which were then in- jected into the nonlinear system through loud- speakers positioned on the right side of the cavity. The output of the physical system was measured using microphones below the meta- scatterers. We investigated the vowel classifi- cation performance of two-layer acoustic-PNN (Fig. 2A). To compare the results of PhyLL with ideal BP and in silico training, we accu- rately modeled the forward pass of acoustic-PPN by a digital neural network (supplementary text, section 2.1). When trained using PhyLL, the acoustic-PNN achieved a classification accuracy of 98.88% and 97.31% for train and test datasets, respectively (Fig. 2, D to F). Figure 2C shows the comparison of the clas- sification results obtained for PhyLL, ideal BP, Input layer Physical layer Topology of PNN Vowel dataset 97.14 % 97.21 % A B 100 y c a r u c c A 90 80 70 60 50 40 0 20 40 60 Epoch 80 Train Test 100 E Positive data Layer-wise training with cos-similarity Negative data 1 k s a M 2 k s a M Mapping representations to labels C y c a r u c c A 98 97 96 95 94 93 92 0 Supervised learning Digit-MNIST dataset 97.19 % 96.36 % Train Test 100 200 300 Epoch Unsupervised learning Supervised learning Unsupervised learning Label Data Label Data Output Output Fashion-MNIST dataset 92.27 % 87.79 % Train Test 100 200 Epoch 300 400 Digit-MNIST dataset 97.59 % 96.51 % Train Test 50 100 Epoch 150 200 D 98 y c a r u c c A 96 94 92 90 88 86 84 82 80 0 F y c a r u c c A 99 98 97 96 95 94 93 92 0 Fig. 4. Optics-PNN. (A) The topology of the optics-PNN consists of a two-layer PNN. Each layer comprises an optics-PNN augmented by trainable matrix multiplication. The right panel shows examples of input encoding for supervised and unsupervised contrastive versions along with the corresponding output on a CCD camera for the digit Mnist dataset. (B to D) The train and test classification accuracy versus training epoch for the vowel (B), digit (C), and fashion Mnist (D) tasks. (E) Schematic of the unsupervised version for PNNs (supplementary text, section S2). (F) The classification accuracy on the training and test sets versus training epoch for the unsupervised contrastive version of PhyLL on the digit Mnist dataset. Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 4 of 7 RES EARCH | R E S E A R C H A R T I C L E A B Input layer Physical layer Topology of PNN Transmission matrix Transmission matrix e t a t s m e t s y S Hard perturb Hard perturbation Time D 100 Perturbation PhyLL C 100 PA-BP 80 Perturbation y c a r u c c a t s e T 60 40 20 0 0 20 40 60 80 Iteration 100 120 140 160 y c a r u c c a t s e T 80 60 40 20 0 0 20 40 60 80 Iteration 100 120 140 160 Fig. 5. Robustness of deep PNNs against unpredictable external perturbations. (A) A deep PNN consists of six layers of optics-PNN augmented by trainable matrix multiplication. The deep PNN is trained on vowel datasets and is currently in the inference phase. (B) Applying hard perturbation by adding Gaussian noise with the mean of m and standard deviation of s to the transmission matrix of MMF. (C and D) A comparison between PA-BP (22) (C) and the proposed PhyLL method (D) is presented, with the focus on their ability to recover the classification accuracy after applying perturbation. and in silico training. A schematic of the afore- mentioned methods is provided above Fig. 2C. The complete comparison between different hardware-based training methods is provided in the supplementary text, section 2.1. As demonstrated in Fig. 2C, in silico training performed poorly, reaching only a maximum vowel classification accuracy of ~50%. When there was a gap between the reality and the sim- ulation of a physical system (called the reality- simulation gap), the accuracy of inference would decrease. By contrast, PhyLL succeeded in accu- rately training the acoustic-PNN, performing similarly to the ideal BP algorithm used as a base- line. The key advantage of PhyLL stems from the execution of both forward passes through the physical hardware rather than simulations. Microwave massively parametrized chaotic cavity with structural nonlinearity In the microwave regime, we leveraged a “struc- tural nonlinearity” such that we could imple- ment nonlinear mathematical operations at low power levels with a linear scattering sys- tem. Our system consisted of a chaotic cavity that was massively parametrized by covering one of its walls with a programmable meta- surface. For each meta-atom and each polari- zation, the programmable metasurface offered two possible local boundary conditions. Our setup is shown in Fig. 3B and further detailed in the supplementary text, section 2.4. Although the setup resembles that recently used to im- plement with high fidelity and in situ repro- grammability desired linear transfer functions for signal differentiation (53) and routing (54), in this case we sought a nonlinear mapping. Hence, we defined the metasurface configu- ration as the input and the transfer function as the output of our mathematical operation. This relation is in general nonlinear because of the mutual coupling between meta-atoms caused by their proximity and, notably, the reverberation (55). We embrace reverbera- tion to maximize the nonlinearity, whereas previous work (50) has sought to limit the reverberation to implement a linear trans- formation with the same input-output defi- nition (see supplementary text, section 2.4, for further discussion). We randomly grouped our programmable metasurface’s 152 degrees of freedom into 40 macropixels because our mathematical opera- tion necessitated 40 input values. We defined our mathematical operation’s outputs as the transfer function intensities at 20 decorrelated frequencies within the bandwidth of operation of the programmable metasurface (400 MHz around 5.2 GHz). Thereby, in addition to the structural nonlinearity, we added a readout nonlinearity by considering the transfer func- tion’s intensity. To flexibly evaluate the pro- posed approach, we learned a digital surrogate forward model of the configuration to trans- fer function intensity mapping (supplemen- tary text, section 2.4). Then, we constructed Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E the three-layer microwave-PNN shown in Fig. 3A and trained it according to the PhyLL. The training converged after roughly 20 epochs, and the achieved classification accuracy on the unseen test data reached 97.31% (Fig. 3, C to E). Optical multimode fiber with readout nonlinearity In the optics example, we used the experimen- tal transfer matrix data of an optical system that comprised an SLM, a scattering medium consisting of a step-index multimode fiber (MMF), and a complementary metal oxide semiconductor (CMOS) camera (Fig. 1D and materials and methods) (56). The data were encoded onto the SLM, and after passing through the MMF, the resulting optical inten- sity was seen on the camera. The physical optical system performed a complex spatial transformation. Although this transformation was linear in the complex domain, the process became nonlinear as a result of the data being encoded onto the phase (SLM) and the sub- sequent measurement of the intensity squared on the camera. We used an optics-PNN to perform classifi- cation tasks on different datasets: vowel, digit, and fashion Mnist (see Fig. 4; table S1; and supplementary text, section 2.5, for further details). The two-layered optics-PNN achieved high classification accuracy on both the vowel and Mnist dataset. For vowel, we obtained 97.21% and 97.14% accuracy on the training and test sets, respectively. Using only the two- layer optics-PNN, the model achieved 97.19% and 96.36% accuracy on the training and test sets of Mnist, respectively. To implement a deeper optics-PNN, we trained it with six layers on the fashion Mnist dataset, achieving train- ing and test accuracies of 92.27% and 87.79%, respectively (see corresponding results in Fig. 4, B to D). In addition, the unsupervised con- trastive learning version of PhyLL is tested on three layers of optics-PNN (Fig. 4E). The mod- el achieved training and test accuracies of 97.59% and 96.51%, respectively (Fig. 4F and table S2). Results were also consistent with a recent preprint for a similar experiment that directly implemented the original forward- forward algorithm, leading to slightly lower efficiency (57). Real-time adaptable learning In this work, we investigated the robustness of PhyLL in the context of real-time and adapt- able learning, where the physical data trans- former may undergo changes as a result of the slow dynamics of the physical system dur- ing the runtime or external hard perturba- tions (see also supplementary text, section 2.8). Let us consider a deep optics–PNN with six layers, as depicted in Fig. 5A, which has al- ready been trained on vowel datasets and is currently in the inference phase. The trans- formation function of each physical system is f0(q), where q is the physical input. We per- turbed the physical systems at a specific time (examples of such perturbations include changes in the MMF state or the positions of lenses or masks, etc.), which results in a change in the transformation function of each physical sys- tem from f0(q) to fp(q) (Fig. 5B). To show this, we perturbed the transmission matrix of the optical setup by adding a Gaussian noise with mean m and standard deviation s. As observed in Fig. 5D, the test accuracy dropped as ex- pected after applying the perturbation. The question now is whether the training meth- od can restore the accuracy by retraining the optics-PNN. We compared our results with the PA-BP method (22), which uses a digital model for the backward pass and the physical system for the forward pass. PA-BP struggled to restore accuracy with increasing perturba- tion intensity (Fig. 5C). For instance, the test accuracy oscillated around 55% for a small per- turbation (red dots in Fig. 5C) and worsened further for more intense perturbations. By contrast, the proposed PhyLL could easily re- cover accuracy after a few epochs, regardless of the intensity of the perturbation applied (Fig. 5D). This adaptability could be attributed to PhyLL executing both forward passes through the physical hardware rather than digital models. By contrast, the PA-BP method relied on a digital model that lost its accuracy when subjected to hard perturbation, necessitating retraining from scratch or comprehensive hy- perparameter tuning. Discussion Because of the unprecedented growth in the size of ANNs, such as large language models (LLMs) that are expected to increase unceas- ingly, the costs of both the training and infer- ence phases of these networks have increased exponentially. Specialized hardware, such as PNNs, have the potential to drastically de- crease these costs. Anderson et al. (21) recently projected an inference–time energy–efficiency advantage of ~8000× compared with that of digital-electronic processors for large-scale future transformer models. The training meth- od proposed in this paper could serve as a viable candidate for training these optical LLMs, potentially offering substantial energy efficiency and speed advantages. We further examine these in the supplementary text, sec- tion 2.10. Implementing large-scale LLMs with optics still faces a few challenges, such as the current SLM capacity limited to a few million parameters—far from the billions required. However, there are no fundamental roadblocks to achieving billion-parameter optical archi- tectures and energy-efficient PNNs. RE FERENCES AND NOTES 1. Y. LeCun, Y. Bengio, G. Hinton, Nature 521, 436–444 (2015). 2. L. Deng, D. Yu, Found. Trends Signal Process. 7, 197–387 (2014). 3. M. Asghari, A. V. Krishnamoorthy, Nat. Photonics 5, 268–270 (2011). 4. D. Patterson et al., Carbon Emissions and Large Neural Network Training arXiv:2104.10350 [cs.LG] (2021). 5. G. Wetzstein et al., Nature 588, 39–47 (2020). 6. X. Lin et al., Science 361, 1004–1008 (2018). 7. Z. Wu, M. Zhou, E. Khoram, B. Liu, Z. Yu, Photon. Res. 8, 46–50 (2020). 8. T. W. Hughes, I. A. D. Williamson, M. Minkov, S. Fan, Sci. Adv. 5, eaay6946 (2019). 9. A. Momeni, R. Fleury, Nat. Commun. 13, 2651 (2022). 10. M. Romera et al., Nature 563, 230–234 (2018). 11. J. Grollier et al., Nat. Electron. 3, 360–370 (2020). 12. T. Chen et al., Nature 577, 341–345 (2020). 13. H.-C. Ruiz Euler et al., Nat. Nanotechnol. 15, 992–998 (2020). 14. Y. Shen et al., Nat. Photonics 11, 441–446 (2017). 15. L. Appeltant et al., Nat. Commun. 2, 468 (2011). 16. A. Momeni, X. Guo, H. Lissek, R. Fleury, Physics-inspired Neuroacoustic Computing Based on Tunable Nonlinear Multiple-scattering arXiv:2304.08380 [cs.SD] (2023). 17. A. Momeni, X. Guo, H. Lissek, R. Fleury, in 2022 Sixteenth International Congress on Artificial Materials for Novel Wave Phenomena (Metamaterials 2022) (IEEE, 2022), pp. 314–316. 18. G. W. Burr et al., Adv. Phys. X 2, 89–124 (2017). 19. D. Marković, A. Mizrahi, D. Querlioz, J. Grollier, Nat. Rev. Phys. 2, 499–510 (2020). 20. M. Prezioso et al., Nature 521, 61–64 (2015). 21. M. G. Anderson, S.-Y. Ma, T. Wang, L. G. Wright, P. L. McMahon, Optical transformers arXiv:2302.10360 [cs.ET] (2023). 22. L. G. Wright et al., Nature 601, 549–555 (2022). 23. M. Nakajima et al., Nat. Commun. 13, 7847 (2022). 24. T. P. Lillicrap, A. Santoro, L. Marris, C. J. Akerman, G. Hinton, Nat. Rev. Neurosci. 21, 335–346 (2020). 25. S. Pai et al., Experimentally realized in situ backpropagation for deep learning in nanophotonic neural networks arXiv:2205. 08501 [cs.ET] (2022). 26. X. Guo, T. D. Barrett, Z. M. Wang, A. Lvovsky, Photon. Res. 9, B71–B80 (2021). 27. T. W. Hughes, M. Minkov, Y. Shi, S. Fan, Optica 5, 864–871 (2018). 28. S. Bandyopadhyay et al., Single chip photonic deep neural network with accelerated training arXiv:2208.01623 [cs.ET] (2022). 29. J. F. Miller, S. L. Harding, G. Tufte, Evol. Intell. 7, 49–67 (2014). 30. J. Bueno et al., Optica 5, 756–760 (2018). 31. G. Hinton, The forward-forward algorithm: Some preliminary investigations arXiv:2212.13345 [cs.LG] (2022). 32. H. Zhu, Y. Chen, G. Hu, S. Yu, Electronics 12, 147 (2022). 33. H.-C. Lee, J. Song, SymBa: Symmetric backpropagation-free contrastive learning with forward-forward algorithm for optimizing convergence arXiv:2303.08418 [cs.CV] (2023). 34. R. F. Srinivasan et al., Forward learning with top-down feedback: Empirical and analytical characterization arXiv:2302. 05440 [cs.LG] (2023). 35. J. R. Movellan, in Connectionist Models, D. S. Touretzky, J. L. Elman, T. J. Sejnowski, G. E. Hinton, Eds. (Elsevier, 1991), pp. 10–17. 36. D.-H. Lee, S. Zhang, A. Fischer, Y. Bengio, in Machine Learning and Knowledge Discovery in Databases: European Conference, ECML PKDD 2015, Proceedings Part I, A. Appice et al., Eds. (Springer, 2015), pp. 498–515. 37. B. Scellier, Y. Bengio, Front. Comput. Neurosci. 11, 24 (2017). 38. A. N. Gomez et al., J. Mach. Learn. Res. 23, 171 (2022). 39. E. Belilovsky, M. Eickenberg, E. Oyallon, in Proceedings of the 36th International Conference on Machine Learning, K. Chaudhuri, R. Salakhutdinov, Eds. (PMLR, 2019), pp. 583–593. 40. Y. Xiong, M. Ren, R. Urtasun, in Advances in Neural Information Processing Systems, H. Larochelle, M. Ranzato, R. Hadsell, M. F. Balcan, H. Lin, Eds. (Curran Associates, Inc., 2020), pp. 11142–11153. 41. S. A. Siddiqui, D. Krueger, Y. LeCun, S. Deny, Blockwise self-supervised learning at scale arXiv:2302.01647 [cs.CV] (2023). 42. C. Gallicchio, A. Micheli, L. Pedrelli, Neurocomputing 268, 87–99 (2017). 43. H. Zhou et al., Light Sci. Appl. 11, 30 (2022). 44. M. W. Matthès, P. del Hougne, J. De Rosny, G. Lerosey, S. M. Popoff, Optica 6, 465–472 (2019). 45. A. Silva et al., Science 343, 160–163 (2014). 46. X. Xu et al., Nature 589, 44–51 (2021). 47. A. Babaee, A. Momeni, A. Abdolali, R. Fleury, Phys. Rev. Appl. 15, 044015 (2021). Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E 48. T. Wang et al., Nat. Commun. 13, 123 (2022). 49. H. Rajabalipanah, A. Momeni, M. Rahmanzadeh, A. Abdolali, R. Fleury, Nanophotonics 11, 1561–1571 (2022). 50. P. del Hougne, G. Lerosey, Phys. Rev. X 8, 041037 (2018). 51. A. Momeni et al., IEEE Trans. Antenn. Propag. 69, 7709–7719 (2021). 52. A. Momeni, M. Safari, A. Abdolali, N. P. Kherani, R. Fleury, Phys. Rev. Appl. 15, 034010 (2021). 53. J. Sol, D. R. Smith, P. Del Hougne, Nat. Commun. 13, 1713 (2022). 54. J. Sol, A. Alhulaymi, A. D. Stone, P. Del Hougne, Sci. Adv. 9, eadf0323 (2023). 55. A. Rabault et al., On the tacit linearity assumption in common cascaded models of RIS-parametrized wireless channels arXiv:2302.04993 [cs.IT] (2023). 56. B. Rahmani et al., Nat. Mach. Intell. 2, 403–410 (2020). 57. I. Oguz et al., Opt. Lett. 48, 5249–5252 (2023). 58. A. Momeni, PhyLL: Backpropagation-free Training of Deep Physical Neural Networks, Github (2023); https://github.com/ MomeniAli/PhyLL. 59. A. Momeni, PhyLL: Adding the figures data, Zenodo (2023); https://doi.org/10.5281/zenodo.10075534. ACKN OWLED GMEN TS A.M. and M.M. thank X. Guo for useful discussions. B.R. acknowledges that all materials pertaining to the optical experiment presented in this manuscript have been sourced from previously published material that is publicly available and were acquired during his tenure at the Laboratory of Applied Photonics Devices at EPFL. This work has no connection to his current employer in any capacity or form. Funding: A.M. and R.F. acknowledge funding from the Swiss National Science Foundation under the Eccellenza grant no. 181232. P.d.H. and R.F. acknowledge funding from the ANR-SNF PRCI program (project “MetaLearn”: ANR-22-CE93-0010-01). Author contributions: A.M. conceived the idea, designed the computational engine, and carried out both the theoretical and numerical simulations as well as a part of the acoustic experiment. B.R. provided the optics data and interpretation of machine learning results. M.M. carried out the acoustic experiment. P.d.H. carried out the microwave experiment. R.F. supervised the project. All authors contributed to the interpretation of the results and the writing of the manuscript. Competing interests: The authors declare no competing interests. Data and materials availability: Materials and methods to evaluate the conclusions in the paper are present in the supplementary materials. All other software and data for running the simulations and experiments are available through Github (58). Data underlying the figures are available through Zenodo (59). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi8474 Materials and Methods Supplementary Text Figs. S1 to S18 Tables S1 to S6 References (60–69) Submitted 22 May 2023; resubmitted 25 September 2023 Accepted 7 November 2023 Published online 23 November 2023 10.1126/science.adi8474 Momeni et al., Science 382, 1297–1303 (2023) 15 December 2023 7 of 7
10.1126_science.adk1111
RES EARCH GLYCOSYLATION Palladium catalysis enables cross-coupling–like SN2-glycosylation of phenols Li-Fan Deng1†, Yingwei Wang2†, Shiyang Xu1, Ao Shen1, Hangping Zhu1, Siyu Zhang1, Xia Zhang1, Dawen Niu1* Despite their importance in life and material sciences, the efficient construction of stereo-defined glycosides remains a challenge. Studies of carbohydrate functions would be advanced if glycosylation methods were as reliable and modular as palladium (Pd)-catalyzed cross-coupling. However, Pd-catalysis excels in forming sp2-hybridized carbon centers whereas glycosylation mostly builds sp3-hybridized C–O linkages. We report a glycosylation platform through Pd-catalyzed SN2 displacement from phenols toward bench-stable, aryl-iodide–containing glycosyl sulfides. The key Pd(II) oxidative addition intermediate diverges from an arylating agent (Csp2 electrophile) to a glycosylating agent (Csp3 electrophile). This method inherits many merits of cross-coupling reactions, including operational simplicity and functional group tolerance. It preserves the SN2 mechanism for various substrates and is amenable to late-stage glycosylation of commercial drugs and natural products. T he utility of glycosides in medicinal chem- istry, materials, and biological science is well appreciated (1, 2) but difficulties in their synthesis by glycosylation pose substantial obstacles to exploring their functions. Both reactants in glycosylation— the glycosyl donors and acceptors—are often structurally complex, and the properties of products are profoundly affected by the ab- solute configuration of glycosidic centers. There- fore, an ideal glycosylation method needs to simultaneously address chemo- and stereo- selectivity issues, two persistent challenges in synthesis. Glycoside synthesis has been propelled by the introduction of new donors and their ac- tivating approaches. Most reported glycosylation methods proceed under (Lewis) acid-promoted conditions, which convert glycosyl donors to (equivalents of) oxocarbenium ions for sub- sequent trapping by acceptors (3, 4). These techniques have been the cornerstones in car- bohydrate synthesis and have allowed for pre- paration of complex structures (5–8). However, controlling or predicting stereoselectivity re- mains nontrivial as the glycosylation mechanism often shifts within the SN1/SN2 continuum (9), depending on the properties of reactants and reaction parameters (10). Issues can also arise when labile donors or harsh activating con- ditions are needed, complicating reaction setup. Among the venues to overcome these obsta- cles, developments include the Yu group (11), which exploits selective gold-alkyne interac- tions; the Jacobsen group (12, 13), which ex- plores mild hydrogen-bond catalysis; the Miller 1State Key Laboratory of Biotherapy and Cancer Center, West China Hospital, and School of Chemical Engineering, Sichuan University, Chengdu, China. 2Department of Nuclear Medicine & Laboratory of Clinical Nuclear Medicine, West China Hospital, Sichuan University, Chengdu, China. *Corresponding author. Email: niudawen@scu.edu.cn †These authors contributed equally to this work. group (14, 15), which harnesses the strong Ca–F bond-forming energy; the Nguyen group (16) that employs Lewis base catalysis; and the Codée (17), Takemoto (18), and Loh groups (19), which utilize halogen-bond catalysis, and others based on transition metal catalysis (20, 21). We have reported radical activation of glycosyl donors (22). Despite this progress, there is still a high demand for methods with a gen- eral scope to prepare stereodefined glycosides in a simple and predictable manner, which continues to fuel mechanistic and methodo- logical advancements. The modularity and reliability inherent in Pd-catalyzed cross-coupling reactions have made them indispensable tools in organic synthesis (23, 24). If O-glycosylation could be as straightforward and robust as Pd-catalyzed cross coupling, the downstream exploration of glycosides would be greatly facilitated. However, Pd-catalysis is typically effective in activating and forging sp2-hybridized carbon centers, whereas glycosidic bonds are mostly sp3-hybridized C–O linkages. Strategies that can bridge this gap and channel the power of Pd-catalyzed cross coupling into the field of glycoside synthesis hold considerable potential. O’Doherty and others have applied Pd-catalyzed allylic substi- tution, followed by alkene (di)hydroxylation, for the de novo synthesis of O-glycosides (25, 26). Here, we report a Pd-catalyzed SN2 glycosyl- ation method that commences with the oxida- tive addition (OA). The utility of this approach is showcased in a general and simple SN2 gly- cosylation of phenols, a prominent challenge in O-glycoside synthesis. Reaction design General and straightforward methods for syn- thesizing stereodefined phenolic O-glycosides are highly valuable but remain nontrivial (14, 27, 28). Phenols—glycosylated or not—are abundant in both naturally occurring and man- made compounds (Fig. 1A) (29). Introducing carbohydrate moieties into phenols has proven to be an effective approach for modifying their physical and biological properties in drug dis- covery endeavors. Glycosylation of phenols (3) is complicated as they exhibit modest nucleo- philicity compared with alcohols under acidic conditions (1 to 2; Fig. 1B). Moreover, phenols are ambident nucleophiles, potentially resulting in either O-glycosylated (4) or C-glycosylated products (5). Pd-catalyzed cross-coupling reactions (Fig. 1C, left cycle) are usually initiated by Pd(0)-mediated OA (6 to 7), followed by ligand exchange (7 to 8) and reductive elimination (8 to 9) to give the desired products (here O- arylated phenols 9). Recognizing the limita- tion of this cycle in activating/building Csp3 centers (30), we designed a strategy (Fig. 1C, right cycle) that uses bench-stable, ortho- iodobiphenyl–substituted sulfides (31, 32) 11 as glycosyl donors. The aryl iodide unit in 11 readily undergoes oxidative addition with Pd(0) catalysts, forming an OA complex 12 that acts as an effective glycosyl (Csp3) elec- trophile, likely driven by its tendency to un- dergo Csp2–S reductive elimination (indicated by dashed lines). Nucleophilic attack to 12 by phenoxides 10 proceeds through a clean and general SN2 mechanism, resulting in in- version of the glycosyl center and genera- tion of 13. As a result of the donor activation mechanism, this glycosylation method exhibits a notable tolerance toward functional groups and al- lows using unprotected glycosyl donors. No O-arylation side products are observed from the process (Fig. 1C), indicating that our ap- proach directs the Pd-containing OA complex (such as 12) to transition from an arylating agent (Csp2) to a glycosylating agent (Csp3), unveiling an unprecedented reactivity. The OA complex 12 behaves uniformly as an SN2 electrophile, from fully oxygenated to fully deoxygenated donors, a rarity in carbohydrate chemistry. This method grants access to either isomer of the phenolic O-glycoside products in a predictable manner, many of which were pre- viously challenging to obtain. The transforma- tion occurs under mildly basic conditions and can be performed as easily as a Pd(0)-catalyzed cross-coupling reaction. Reaction validation and condition optimization Our study commenced with the model reaction between sulfide 14 or 15 and 4-methoxy phenol (16) to make O-glycoside 17 or 18 (Fig. 2). The stereoselective synthesis of 2-deoxyglycosides of phenols (e.g., 17/18) has been difficult due to the lack of a C2-substituent as a stereodirect- ing auxiliary and the susceptibility of the 2-deoxyglycoside products to acid-promoted hydrolysis. Guided by the reaction design in Fig. 1C, we established conditions to make Deng et al., Science 382, 928–935 (2023) 24 November 2023 1 of 8 RES EARCH | R E S E A R C H A R T I C L E A Ph O O HO HO O O OH O HO HO O O OMe OH O R O O N N OH OMe O O HO O O HN O O O NH HN S HO O OH OH O N N R = Me, Scrophuloside C R = OH, Scrophuloside D Glycosylated Genistein Afeletecan (Phase I, anticancer) HO RO O LG RO O 3 RO O O + RO 1 or equivalents 2 4 O 5 OH B C O 9 = oIB This Work Csp3 N2) I 6 Pd(0)-L O I S RO 11 Classical Csp2 Csp3 ? O Pd L 8 L I Pd S O I Pd L RO 7 12 O RO 10 O O 13 Fig. 1. Glycosylated phenols: background and synthetic approaches. (A) Representative examples of glycosylated phenols. (B) Glycosylation of phenols: challenges and limitations. (C) Comparison of Pd-catalyzed C–O cross-coupling and Pd-catalyzed SN2 glycosylation (this work). FG, functional group. O-glycosides 17/18 from 14/15 in high yields, with dibenzothiophene (19) formed as a byproduct. The b-O-glycoside 17 was obtained from the a-S-glycoside 14, and a-O-glycoside 18 formed if b-S-glycoside 15 was employed. The clean inversion of the glycosidic centers in 14/15 suggests that this Pd-catalyzed glycosylation pro- ceeds by means of an SN2-type mechanism. Gen- eral and operationally simple SN2-glycosylation methods remain rare, despite their high value as tools (12, 34–36) to access stereodefined glycosides. Even scarcer are methods that can afford both stereoisomers, as accessing the cor- responding donors with defined and stable con- figurations is not always simple. In our case, sulfide donors (with general structure 11) could be prepared from the corresponding 1-glycosyl thioacetates in one pot at decagram scales (see SM section 3) and stored for months without precautions to avoid air or moisture. Employ- ing the Pd(0)-mediated OA as a donor-activating Deng et al., Science 382, 928–935 (2023) 24 November 2023 2 of 8 RES EARCH | R E S E A R C H A R T I C L E OBn O BnO BnO SoIB 14 (1.0 equiv.) OBn O SoIB BnO BnO 15 (1.0 equiv.) OBn O O BnO BnO Pd(0)-Xantphos (2 mol%) K2CO3 (2.0 equiv.) OMe 17 (95% yield, : < 1:19) HO OMe 16 (1.5 equiv.) BnO BnO OBn O O 18 (90% yield, : > 19:1) OMe Entry Divation from standard Conditions (with 15 used) Conversion Yield ( : ) none no Pd(0)-Xantphos no K2CO3 100% 92% (>19:1) 0% 7% 0% 5% (>19:1) Cs2CO3 instead of K2CO3 100% 94% (>19:1) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 TEA instead of K2CO3 TMG instead of K2CO3 K2CO3 aq. room temperature MeCN instead of Toluene no XantPhos NiXantPhos instead of XantPhos DavePhos instead of XantPhos dppb instead of XantPhos Addition of TEMPO 45% 79% 87% 21% 41% 77% 100% 67% 100% 100% Ligands O H N O PPh2 PPh2 PPh2 PPh2 PCy2 Ph2P PPh2 4 NMe2 Xantphos Ni-Xantphos Davephos dppb Fig. 2. Reaction validation and condition optimization. Reactions in this table were performed at 0.05 mmol scale at 0.1 M concentration for 24 hours. Yields and conversions are determined by 1H NMR analysis using 1,3,5- trimethoxy benzene as an internal standard. Diastereomeric ratios were determined by NMR analysis of crude reaction mixture. TMG, 1,1,3,3-tetramethylguanidine; TEMPO, (2,2,6,6-tetramethylpiperidin-1-yl)oxyl; Bn, benzyl. approach, this method provides either isomer of the phenolic O-glycosides. The reaction proceeds at 60°C and only requires K2CO3 as base and a catalytic amount (2 mol%) of Pd(0) catalyst as- sembled from Pd(dba)2 and Xantphos. We conducted control experiments to iden- tify factors influencing the performance of this method. Little reaction occurred in the absence of Pd(0) catalyst (entry 2) or a suitable base (entry 3). Inorganic bases K2CO3 (entry 1) and Cs2CO3 (entry 4) exhibited better effects than organic bases such as Et3N (entry 5) and TMG (entry 6). A saturated aqueous solution of K2CO3 could be used (entry 7), suggesting that the re- action has a good tolerance to water. When the reaction is conducted at room temperature 20% (>19:1) 54% (2:3) 71% (9:1) 19% (>19:1) 40% (1:1) 75% (>19:1) 85% (4:1) 56% (>19:1) 90% (3:1) 91% (10:1) Byproduct S 19 (roughly 25°C), the conversion is relatively low within the same time period (entry 8). The use of more polar solvents such as MeCN greatly eroded the stereochemical purity of products (entry 9). The added ligands exert pronounced effects on the reaction efficiency (entries 10 to 13), likely through modulating the properties of OA complex such as 12 in Fig. 1C. Unlike phosphine-based ligands, nitrogen-based ligands and N-heterocyclic carbene ligands we tried were less effective (see SM, section 4). Addition of a stoichiometric amount of TEMPO was fully tolerated, essentially excluding a radical-based mechanism (entry 14) (22, 37). Substrate scope It is rare for a glycosylation method to hold high efficiency for a broad array of substrates as the reaction mechanism (and outcome) often varies with the reactivities of donors and ac- ceptors (38–41). Our method accommodates diverse glycosyl units (Fig. 3), providing ei- ther of the two possible stereoisomers with high purities (22a-m). For example, donors bearing benzyl (22a, 22k, 22l), acetyl (22b, 22f), benzylidene (22d) and silyl–protecting groups (22c, 22e) could all be used. Various other 2-deoxypyranosyl groups (22f-j) were in- stalled with similar efficiencies. Our meth- od could be adopted to construct the more electron-rich furanosyl linkages: both iso- mers of 2-deoxyribosides were generated cleanly (22k). This method is not limited to 2-deoxy sugars, and the SN2 mechanism op- erates with fully oxygenated (22l and 38g/h in Fig. 4) and fully deoxygenated tetrahydro- pyranyl (22m) donors. 2-O-acetyl or 2-N-acetyl protected donors were unsuccessful, likely be- cause of interference from these neighbor- ing participating groups (fig. S6). Aliphatic alcohols are almost inert under the current conditions, allowing the use of unprotected glycosyl donors, as shown by examples 22g-j. The results attest to the exceptional func- tional group tolerance of this OA-initiated glycosylation method. It is worth highlight- ing that many of the glycosyl units in Fig. 3 are deoxygenated and electron rich. To obtain the corresponding O-glycosides with high stereochemical purities would be tedious by conventional methods, due to dearth of suitable donors, lack of stereo-directing auxil- iaries, and susceptibility of products to acid- promoted hydrolysis. Both electron-rich (22w, 22ae, 22ah) and electron-deficient phenols (22n-r) were ac- commodated, with no product arising from C-glycosylation observed. Phenols bearing an ortho-substituent were competent substrates (22u-w, 22y-z). Functional groups such as aldehydes (22q), esters (22r), secondary amides (22af), nitriles (22p), and ketones (22x) were tolerated. The terminal alkene group in 22v did not isomerize to conjugate with a phenyl ring. Deng et al., Science 382, 928–935 (2023) 24 November 2023 3 of 8 RES EARCH | R E S E A R C H A R T I C L E O SoIB HO Pd(0)-Xantphos (2 mol%) O O RO + 20 N N 21 K2CO3 (2.0 equiv.) Toluene (0.1 M), 60 °C RO N 22 Scope of donor, acceptor = 16 or 4-CO2MePhOH OBn O BnO BnO OBn O BnO BnO OAc O AcO AcO OAc O AcO AcO OTBS O HO HO OTBS O HO HO Ph O O HO Ph O O O HO O 22a- 90% ( : > 19:1) 22a- 95% ( : < 1:19) 22b- 88% ( : = 8:1)a 22b- 71% ( : = 1:13)a 22c- 57% ( : > 19:1) 22c- 70% ( : = 1:17) 22d- 78% ( : > 19:1) 22d- 63% ( : = 1:13) tBu tBu tBu Si O O HO tBu O Si O O HO O AcO OAc O AcO AcO OAc O AcO Me HO O OH Me HO O OH Me O Me O OH HO OH HO 22e- 81% ( : = 15:1) 22e- 69% ( : = 1:10) 22f- 89% ( : = 5:1)a 22f- 78% ( : = 1:8)a 22g- 67% ( : = 5:1) 22g- 72% ( : = 1:16) 22h- 68% ( : = 6:1) 22h- 77% ( : = 1:10) HO HO O HO HO O O OH HO O HOOH BnO O BnO O OBn OBn 22i- 84% ( : = 8:1) 22i- 51% ( : = 1:10) 22j- 77% ( : = 10:1) 22j- 73% ( : = 1:6) 22k- 56% ( : = 6:1) 22k- 77% ( : = 1:10) BnO BnO OBn O BnO 22l- 71% ( : > 19:1)a 22l- 86% ( : < 1:19)a Scope of acceptor, donor = 14/15 O O O O O O Br CF3 CN CHO CO2Me Bpin O 22m 92% (95:5 er) O O O 22n- 96% ( : > 19:1) 22n- 83% ( : = 1:14) Br O 22u- 82% ( : > 19:1) 22u- 83% ( : < 1:19) O N 22ab- 85% ( : > 19:1) 22ab- 72% ( : < 1:19) 22o- 87% ( : > 19:1) 22o- 88% ( : < 1:19) 22p- 93% ( : = 15:1)b 22p- 88% ( : < 1:19)b 22q- 73% ( : = 10:1)b 22q- 68% ( : = 1:10)b O O OMe O Me O 22w- 72% ( : > 19:1)c 22w- 80% ( : = 1:16)c 22x- 95% ( : > 19:1) 22x- 74% ( : < 1:19) 22r- 94% ( : > 19:1) 22r- 73% ( : = 1:11) O F Br Me 22y- 95% ( : > 19:1) 22y- 84% ( : < 1:19) 22v- 76% ( : = 7:1) 22v- 85% ( : < 1:19) O N Cl F 22ac- 88% ( : > 19:1) 22ac- 65% ( : = 1:16) O N O 22ad- 72% ( : > 19:1) 22ad- 80% ( : < 1:19) O O O N BocHN CO2Me 22ae- 69% ( : > 19:1) 22ae- 85% ( : = 1:17) 22af- 90% ( : > 19:1) 22af- 95% ( : < 1:19) 22ag- 62% ( : = 10:1)c 22ag- 70% ( : < 1:19)c 22ah- 80% ( : = 10:1) 22ah- 78% ( : = 1:14) 22s- 81% ( : > 19:1) 22s- 67% ( : < 1:19) 22t- 96% ( : > 19:1) 22t- 88% ( : < 1:19) O Cl Cl O N S 22z- 92% ( : > 19:1) 22z- 89% ( : < 1:19) 22aa- 79% ( : > 19:1) 22aa- 88% ( : < 1:19) Me Me O O O OMe OMe OMe Fig. 3. Substrate scope. Unless otherwise noted, reactions in this table were performed at 0.1 or 0.2 mmol scale in toluene (0.1 M) for 24 hours, using Pd(PPh3)4 (2 mol%), Xantphos (4 mol%), and K2CO3. Isolated yields are reported. Diastereomeric ratios were determined by NMR analysis of crude reaction mixture. The areaction was run at 80°C; bNi-XantPhos was used instead of XantPhos; cCs2CO3 as base. See SM, section 5 for experimental details. Potentially chelating heterocycles including dioxolanes (22t), thiazoles (22aa), quinolines (22ab), pyridines (22ac), oxazoles (22ad), and morpholines (22ae) were incorporated. Protic hydrogen atoms in secondary amides (22af) were compatible. Tyrosine derivatives could be glycosylated (22af), and the a-stereocenter in the amino acid backbone stayed intact. Aryl bromides/chlorides (22n, 22u, 22y-z) did not interfere with this method, likely because the oxidative addition to aryl iodide units in donor 14 or 15 is a faster process. The remain- ing aryl halide groups could serve as handles for further derivatizations (see below). Partic- ularly noteworthy is that aryl boronic esters (22s) survived the reaction conditions with- out undergoing a Suzuki-Miyaura reaction, highlighting the distinctive reactivity of our Deng et al., Science 382, 928–935 (2023) 24 November 2023 4 of 8 RES EARCH | R E S E A R C H A R T I C L E A OBn O BnO BnO OBn O BnO BnO O O Ph O OH O O O BnO BnO OBn O O OBn O O BnO BnO MeO CHO S N Me 23 from Chrysin 52% ( : > 19:1) 24 from 2-Hydroxy-anthraquinone 72% ( : = 6:1) 25 from Vanillin 70% ( : > 19:1) 26 from Rotigotine 73% ( : < 1:19) Ph O O HO O O OTBS O O HO HO H H H OMe H H H BnO BnO Me OH OBn O O MeO Me Me H N O BnO BnO OBn O O MeO BnO BnO OBn O O Ph Me N iPr iPr 27 from Tolterodine 69% ( : < 1:19) OBn O BnO BnO O O O 28 from Estrone 66% ( : > 19:1) 29 from Ethynylestradiol 73% ( : = 1:14)a 30 from Capsaicin 74% ( : = 1:19) 31 from Eugenol 78% ( : < 1:19) 32 from 7-Hydroxycoumarin 82% ( : < 1:19) OBn BnO BnO O O OH Me N Me OTBS O O HO HO Cl O Cl Cl OBn O O BnO BnO O O OBn O O BnO BnO Me OMe Me O O O N F N O F OH 36 from Ezetimibe 74% ( : < 1:19) 33 from O-desmethylvenlafaxine 86% ( : < 1:19) 34 from Triclosan 76% ( : = 1:18)a 35 from Mycophenolate mofetil 69% ( : = 1:5) BnO BnO OBn O Me Me O O OMe O OH O O O BnO OBn OBn Me O N N TESO O O Me 37 from Icaritin 68% ( : > 19:1) SN-38 derivatives = BnO BnO OBn O OBn O BnO BnO BnO OBn O BnO BnO OBn O BnO 38a 64% ( : < 1:19) 38b 75% ( : > 19:1) 38c 84% ( : = 1:11) 38d 64% ( : > 19:1) BnO O BnO O OBn OBn Me O OBn BnO OBn O OBn OBn BnO 38e 60% ( : < 1:19) 38f 74% ( : > 10:1) 38g 51% ( : > 19:1)b 38h 50% ( : < 1:19)b B BnO BnO OBn O O 40 35% yield OBn O O BnO BnO 42 55% yield CO2Me NHBoc One pot, multistep reactions Pd(0)-Xantphos K2CO3 60 to 100 °C Glycosylation/ Suzuki-Miyaura OBn O SoIB BnO BnO HO Br 14 or 15 (1.0 equiv.) + 39 (1.5 equiv.) Me Glycosylation/ Buchwald-Hartwig Bpin H2N H N O Pd(0)-Brettphos K2CO3 60 to 100 °C the 3rd coupling partner (2.0 equiv.) Pd(0)-Xantphos K2CO3 60 to 100 °C Glycosylation/ Suzuki-Miyaura Glycosylation/ Sonogashira Pd(0)-Xantphos K2CO3 60 to 100 °C BnO BnO OBn O O 41 from Loratadine 50% yield N N OEt O OBn O BnO BnO H H O MeH O Me HO 43 from ethisterone 61% yield Fig. 4. Synthetic application. (A) Glycosylation of natural products and commercial drugs. (B) One-pot, multistep, multicomponent reactions. Reactions in this table were performed at 0.1 or 0.2 mmol scale in toluene (0.1 M) for 24 hours, using Pd(PPh3)4 (2 mol%), Xantphos (4 mol%), and Cs2CO3. Isolated yields are reported. Diastereomeric ratios were determined by NMR analysis. aK2CO3 used as base; bReaction performed at 100 °C. See SM sections 6 and 7 for experimental details. Pd(II) OA complexes (i.e., 12 in Fig. 1C) as glycosyl (Csp3) electrophiles. Synthetic application To demonstrate the utility of this method, we applied it in the modification of natural prod- ucts and commercial drugs (Fig. 4A). Simple phenols such as vanillin (25) and triclosan (34) were glycosylated smoothly. Internal alkene groups are accommodated, as exemplified by the formation of 30-31 and 35. The tertiary amine groups in drug molecules may inter- fere with acid-promoted glycosylation meth- ods, but show excellent compatibility with our conditions (26-27, 33). Under our conditions, protection of aliphatic alcohols is unnecessary (29, 33). In the case of chrysin (23), the C7- OH is glycosylated with high regioselectivity. Deng et al., Science 382, 928–935 (2023) 24 November 2023 5 of 8 RES EARCH | R E S E A R C H A R T I C L E OAc O S I 44 A AcO AcO B Pd(PPh3)4 CH2Cl2 71% yield OAc AcO AcO O I Ph3P Pd S TS-I AcO AcO OAc O I Ph3P Pd 45 1 S 1.87 Å = X-Ray structure CCDC 2283248 (H atoms omitted) 1.39 Å 1 + OH OMe 16 K2CO3 Toluene AcO AcO OAc O O OMe 22b Without Ligand 70% yield, : = 1:5 XantPhos 73% yield, : = 1:11 AcO AcO 23.3 31.5 11.2 B3LYP-D3/def2-tzvp/SMD(toluene)//B3LYP-D3/def2-svp G (kcal/mol) OAc O S Int-I 46 -10.7 Ph3P S Pd TS-III -3.6 7.1 Ph3P S Pd + I- + OAc O O AcO AcO 47 Int-II -20.8 Ph3P Pd + S 19 Internal competition (both 3 and 48 are added) External competition (only 48 is added) ln krel CO2Me conversion (%) of 15 @ 90 min CF3 Br of R OMe H Me R R = OMe Me H Br CF3 w/o 48 45 + 0 O 46 OAc O O AcO AcO I Ph3P Pd S C OH 3 OH R 48 + 15 Pd(PPh3)4 (2 mol%) Xantphos (4 mol%) K2CO3 (2.0 equiv.) Toluene (0.1 M) TS-II OBn O O 49 + OBn O O 50 BnO BnO BnO BnO Fig. 5. Mechanistic studies. (A) OA complex 45, formation, isolation, structure and reactivity. (B) Computed energetics for potential reaction pathways. Free energies are computed by B3LYP-D3/def2-tzvp/SMD(toluene)//B3LYP-D3/def2-svp. (C) Reactivity of various phenols. Procedures to determine relative rate krel: phenol (0.1 mmol), substituted phenol 48 (0.1 mmol) and glycosyl donor 15 (0.1 mmol) are allowed to react under standard conditions and stopped at ~30% conversion. krel, conversion of substituted phenol divided by conversion of phenol. The results are the average of three runs. Error bars represent standard deviations. See SM section 8 for experimental details. For polyphenols such as icaritin (37), two gly- cosyl units can be installed at a time, both with high selectivities. Glycosylated anthraquinones are ubiquitous in nature and we show that 2-hydroxylanthraquinone is a competent sub- strate in our reaction (24). O-glycosylated cou- marins are frequently employed as fluorescent probes for detection of glycosidases (42), and coumarins are glycosylated cleanly (32). Ster- iodal phenols such as estrone (28) and estradiol (29) are modified. A glycosyl unit was installed onto ezetimibe (36), a cholesterol absorption Deng et al., Science 382, 928–935 (2023) 24 November 2023 6 of 8 RES EARCH | R E S E A R C H A R T I C L E inhibitor containing a sensitive b-lactam unit. SN-38 is a potent anticancer agent derived from camptothecin (43) but is poorly soluble in wa- ter. Conjugation with a sugar may improve the pharmacological profile of the parent molecule; see Afeletecan in Fig. 1A. We mounted an array of glycosyl units, including both C2-deoxygenated (38a-f) and C2-oxygenated sugars (38g-h), onto an SN-38 derivative (Fig. 4A). Pd-catalyzed SN2 glycosylation could be performed in tandem with other Pd-catalyzed cross-coupling reactions, affording carbohydrate- containing compounds by one-pot, multistep processes in which a single Pd(0)-complex catalyzes two distinct steps (Fig. 4B). Mixing glycosyl donor 14 or 15, m-bromo phenol (39), and aryl boronic ester in one pot with a Pd(0)- catalyst and a base, the Pd-catalyzed glycosyla- tion and the Suzuki-Miyaura coupling proceed in sequence, to give loratadine derivative 41 or L-tyrosine derivative 40 in useful overall yields. The glycosylation could also be merged with Buchwald-Hartwig (44, 45) coupling if the ligand is switched to Brettphos (46) and m-toluamide as the third coupling partner; consecutive formation of Csp3–O and Csp2–N bonds delivered 42 smoothly. Lastly, a sequence composed of glycosylation/Sonogashira cou- pling provides ethisterone-sugar conjugate 43 in 61% overall yield. These examples il- lustrate selectivity of our Pd-catalyzed glyco- sylation and suggest a rapid approach to make glycoconjugates. Mechanistic studies Acetyl-protected sulfide donor 44 reacted with Pd(PPh3)4 smoothly, and the OA complex 45 could be isolated from the mixture by column chromatography (Fig. 5A). The facility of the OA step may be attributable to the sulfur atom in 44, which could pre-coordinate with Pd(0). Complex 45 is a crystalline solid and its mo- lecular structure was verified by x-ray crystal- lography. Similar to other classical Pd(II) OA complexes (47), the Pd center in 45 adopts a (slightly twisted) square planar configuration, with the two neutral ligands (i.e., sulfur and phosphine atoms) occupying para positions. From the solid-state structure of 45, we noted that the C1–S bond is slightly elongated (1.83 to 1.87 Å) and the C1–O bond is shortened (1.41 to 1.39 Å) from their normal lengths (48), indicating buildup of a positive charge at the C1 position. Treating 45 with 4-methoxy phenol (16) in the presence of K2CO3 afforded 22b with de- cent efficiency and inverted configuration. Ad- dition of external ligand Xantphos gave similar results. A small amount of 45 (2 mol%) could catalyze the reaction between 44 and 16 to form 22b. These results support 45 as a re- active intermediate in our process. We next performed DFT calculations (B3LYP- D3/def2-tzvp/SMD(toluene)//B3LYP-D3/def2-svp) employing 45 and phenoxide 46 (or cesium phenoxide, see SM section 9) as model sub- strates (Fig. 5B). Two potential pathways were examined: In principle, reductive elimination of C–S bond in 45 to yield sulfonium Int-I (by means of TS-I), followed by phenoxide attack from 46 could afford glycoside 47. Alterna- tively, direct attack of phenoxide 46 toward 45 by means of TS-II, followed by reductive elimination of the C–S bond (45) in Int-II would provide 47 as well, affording 19 as a byproduct and regenerating Pd(0) catalyst. Upon comparing the free energies of species TS-I and TS-II (note: with different charge states), we observed that the pathway through TS-II exhibits lower barriers. Presumably, co- ordination of the sulfur atom to the Pd center polarized the C–S bond in 45 and made it electrophilic enough toward phenoxide attack. We also considered a scenario where the iodine anion dissociates early from 45 before SN2 dis- placement by 46 (see SM, section 9). We also compared the relative reactivities of various phenols bearing different para- substituents (48) in our glycosylation reaction (Fig. 5C). In internal competition experiments, those with electron-withdrawing substituents (49) react at faster rates, presumably because they are more easily deprotonated under basic conditions. By external competition experi- ments, we found that the turnover frequency, as inferred from the conversion of 15, increases with the electron-withdrawing ability of the para-substituent. The absence of phenol 48 resulted in essentially no reaction. These re- sults indicate the involvement of phenoxide nucleophiles in the turnover-determining step and in turn lend some further support for the pathway through TS-II. Although addi- tional experiments are warranted to elucidate mechanistic details, the ability of Pd-containing OA complex to serve as a glycosyl (Csp3) elec- trophile was quite general (see examples in Figs. 3 and 4). Conclusions We developed a strategy that exploits Pd(0)- mediated oxidative addition—the initial step in classical cross-coupling reactions—as a tool to activate glycosyl donors. The key to this success was the use of aryl iodide-containing glycosyl sulfides as donors, which upon re- action with Pd(0)-catalysts furnished Pd(II)- containing OA complexes that act as glycosyl (Csp3) electrophiles. The following glycosidic bond-forming stage caused clean inversion of the glycosyl centers in donors, and either ste- reoisomer could be obtained in a predictable manner. This approach enabled a general meth- od for SN2-glycosylation of phenols, allowing for the synthesis of phenolic O-glycosides that were previously challenging to access. The mechanism grants this reaction oper- ational simplicity and functional group tol- erance. No acid or cryogenic conditions are required, and the reaction can be set up sim- ilarly to other Pd-catalyzed C–O cross-coupling reactions. Moreover, the method is amenable to late-stage glycosylation of a wide range of commercial drugs and natural products. The generality and mildness of the method is fur- ther showcased in several one-pot, multistep, multicomponent reactions. We anticipate that this study will bring opportunities in Pd-mediated glycosylation reactions, enabling advancements in carbohydrate synthesis and its application in various fields. REFERENCES AND NOTES 1. A. Varki et al., Essentials of Glycobiology (Cold Spring Harbor, 2. 2022), ed. 4 J. J. Barchi, Comprehensive Glycoscience: From Chemistry to Systems Biology (Elsevier, 2021), ed. 2 3. A. V. Demchenko, Handbook of Chemical Glycosylation. (Wiley, 2008). 4. D. B. Werz, S. Vidal, Modern Synthetic Methods in Carbohydrate Chemistry: From Monosaccharides to Complex Glycoconjugates (Wiley, 2014). 5. S. J. Danishefsky, Y.-K. Shue, M. N. Chang, C.-H. Wong, Acc. Chem. Res. 48, 643–652 (2015). 6. X. Zou et al., J. Am. Chem. Soc. 140, 3120–3127 (2018). 7. X. Xiao et al., J. Am. Chem. Soc. 142, 5498–5503 (2020). 8. W. Yao et al., Nat. Synth. 1, 854–863 (2022). 9. P. O. Adero, H. Amarasekara, P. Wen, L. Bohé, D. Crich, Chem. Rev. 118, 8242–8284 (2018). 10. S. Chatterjee, S. Moon, F. Hentschel, K. Gilmore, P. H. Seeberger, J. Am. Chem. Soc. 140, 11942–11953 (2018). 11. B. Yu, Acc. Chem. Res. 51, 507–516 (2018). 12. Y. Park et al., Science 355, 162–166 (2017). 13. Q. Li, S. M. Levi, C. C. Wagen, A. E. Wendlandt, E. N. Jacobsen, Nature 608, 74–79 (2022). 14. T. J. Wadzinski et al., Nat. Chem. 10, 644–652 (2018). 15. G. Pelletier, A. Zwicker, C. L. Allen, A. Schepartz, S. J. Miller, J. Am. Chem. Soc. 138, 3175–3182 (2016). 16. J. Li, H. M. Nguyen, Acc. Chem. Res. 55, 3738–3751 (2022). 17. R. Castelli et al., Chem. Asian J. 9, 2095–2098 (2014). 18. Y. Kobayashi, Y. Nakatsuji, S. Li, S. Tsuzuki, Y. Takemoto, Angew. Chem. Int. Ed. 57, 3646–3650 (2018). 19. C. C. J. Loh, Nat. Rev. Chem. 5, 792–815 (2021). 20. M. J. McKay, H. M. Nguyen, ACS Catal. 2, 1563–1595 (2012). 21. X. Li, J. Zhu, Eur. J. Org. Chem. 2016, 4724–4767 (2016). 22. W. Shang, D. Niu, Acc. Chem. Res. 56, 2473–2488 (2023). 23. C. C. C. Johansson Seechurn, M. O. Kitching, T. J. Colacot, V. Snieckus, Angew. Chem. Int. Ed. 51, 5062–5085 (2012). 24. K. C. Nicolaou, P. G. Bulger, D. Sarlah, Angew. Chem. Int. Ed. 44, 4442–4489 (2005). 25. D. Ray, G. A. O’Doherty, in Protecting Groups: Strategies and Applications in Carbohydrate Chemistry, Edited by S. Vidal (Wiley, 2019), Ch. 9. 26. R. S. Babu, G. A. O’Doherty, J. Am. Chem. Soc. 125, 12406–12407 (2003). 27. M. Jacobsson, J. Malmberg, U. Ellervik, Carbohydr. Res. 341, 1266–1281 (2006). 28. Y. Li, H. Mo, G. Lian, B. Yu, Carbohydr. Res. 363, 14–22 (2012). 29. Z. Rappoport, The Chemistry of Phenols (Wiley, 2003). 30. N. Kambe, T. Iwasaki, J. Terao, Chem. Soc. Rev. 40, 4937–4947 (2011). 31. T. Ooi, M. Furuya, D. Sakai, K. Maruoka, Adv. Synth. Catal. 343, 166–168 (2001). 32. Y. Wang, L.-F. Deng, X. Zhang, Z.-D. Mou, D. Niu, Angew. Chem. Int. Ed. 60, 2155–2159 (2021). 33. C. S. Bennett, M. C. Galan, Chem. Rev. 118, 7931–7985 (2018). 34. X. Ma et al., J. Am. Chem. Soc. 143, 11908–11913 (2021). 35. J. P. Issa, C. S. Bennett, J. Am. Chem. Soc. 136, 5740–5744 (2014). 36. M. H. Zhuo, D. J. Wilbur, E. E. Kwan, C. S. Bennett, J. Am. Chem. Soc. 141, 16743–16754 (2019). 37. Y. Jiang, Y. Zhang, B. C. Lee, M. J. Koh, Angew. Chem. Int. Ed. 62, e202305138 (2023). 38. H. D. Premathilake, A. V. Demchenko, Top. Curr. Chem. 301, 189–221 (2011). 39. K. M. Koeller, C. H. Wong, Chem. Rev. 100, 4465–4494 (2000). Deng et al., Science 382, 928–935 (2023) 24 November 2023 7 of 8 RES EARCH | R E S E A R C H A R T I C L E 40. C. W. Chang et al., Angew. Chem. Int. Ed. 60, 12413–12423 (2021). 41. S. van der Vorm et al., Angew. Chem. Int. Ed. 57, 8240–8244 (2018). 42. B. Thomas et al., Chem. Soc. Rev. 49, 593–641 (2020). 43. F. Li, T. Jiang, Q. Li, X. Ling, Am. J. Cancer Res. 7, 2350–2394 (2017). 44. P. Ruiz-Castillo, S. L. Buchwald, Chem. Rev. 116, 12564–12649 (2016). 45. J. F. Hartwig, Acc. Chem. Res. 41, 1534–1544 (2008). 46. B. P. Fors, D. A. Watson, M. R. Biscoe, S. L. Buchwald, J. Am. Chem. Soc. 130, 13552–13554 (2008). 47. C. Bolm, K. Wenz, G. Raabe, J. Organomet. Chem. 662, 23–33 (2002). 48. H. Liu et al., J. Am. Chem. Soc. 145, 3682–3695 (2023). 49. C. Hansch, A. Leo, R. W. Taft, Chem. Rev. 91, 165–195 (1991). ACKN OWLED GMEN TS We acknowledge P. Yu (SusTEC) for helpful discussions and computational resources. We also thank J. Yang (SCU) for helpful suggestions. Funding: D.N. is supported by National Natural Science Foundation of China (T2221004 and 21922106). Author contributions: D.N. conceived the idea, guided the project, and wrote the manuscript with feedbacks from other authors. L.-F.D. and Y.W. made the initial observations and analyzed the results. L.-F.D., Y.W., A.S., S.Z., and H.Z. explored substrate scope. L.-F.D. and Y.W. performed the mechanistic studies. Under the guidance of X.Z. and D.N, S.X. performed DFT calculations. Competing interests: The authors declare no competing interests. Data and materials availability: Crystallographic data for compound 45 are available from the Cambridge Crystallographic Data Center under reference number CCDC 2283248. All other data discussed in this paper are available in the main text or Supplementary Materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https:// www.sciencemag.org/about/science-licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adk1111 Materials and Methods Supplementary Text Figs. S1 to S9 References (50–63) Submitted 4 August 2023; accepted 24 October 2023 10.1126/science.adk1111 Deng et al., Science 382, 928–935 (2023) 24 November 2023 8 of 8
10.1126_science.adj5328
RES EARCH PHYSICAL CHEMISTRY Direct observation of chirality-induced spin selectivity in electron donor–acceptor molecules Hannah J. Eckvahl1†, Nikolai A. Tcyrulnikov1†, Alessandro Chiesa2†, Jillian M. Bradley1, Ryan M. Young1, Stefano Carretta2*, Matthew D. Krzyaniak1*, Michael R. Wasielewski1* The role of chirality in determining the spin dynamics of photoinduced electron transfer in donor- acceptor molecules remains an open question. Although chirality-induced spin selectivity (CISS) has been demonstrated in molecules bound to substrates, experimental information about whether this process influences spin dynamics in the molecules themselves is lacking. Here we used time-resolved electron paramagnetic resonance spectroscopy to show that CISS strongly influences the spin dynamics of isolated covalent donor–chiral bridge–acceptor (D-Bc-A) molecules in which selective photoexcitation of D is followed by two rapid, sequential electron-transfer events to yield D(cid:129)+-Bc-A(cid:129)–. Exploiting this phenomenon affords the possibility of using chiral molecular building blocks to control electron spin states in quantum information applications. M olecules offer a wide variety of quan- tum properties that could potentially be exploited in qubit architectures for quantum information science (QIS) (1, 2). Moreover, molecules afford the ability to tailor these properties as applications dictate while controlling structure with atomic precision. One such property of growing in- terest is molecular chirality, which plays an essential role in many chemical reactions and nearly all biological processes. Naaman, Waldeck, and coworkers presented evidence of the rela- tionship between molecular chirality and electron spin (3, 4) when they observed that electrons photoemitted from a gold surface coated with a thin film of DNA have a preferred spin state, a phenomenon now known as chirality-induced spin selectivity (CISS) (5). Subsequent experi- ments with molecules bound to metallic, semi- conductor, or magnetic substrates have confirmed a connection between electron motion and spin projection along the chiral axis, which is se- lected to be parallel or antiparallel to the motion depending on the handedness of the chiral molecule (5–9). The spin selectivity of the ef- fect can be very high, even at room tempera- ture, and its theoretical foundations are still being explored (10–17). However, a key prob- lem hindering a fundamental understand- ing of CISS is that it is difficult to separate the role of the substrate from that of the chi- ral molecule. Hence, it is crucial to investigate how CISS affects electron spin dynamics in molecules undergoing electron transfer that are not bound 1Department of Chemistry, Center for Molecular Quantum Transduction and Paula M. Trienens Institute for Sustainability and Energy, Northwestern University, Evanston, IL 60208-3113, USA. 2Università di Parma, Dipartimento di Scienze Matematiche, Fisiche e Informatiche, I-43124 Parma, Italy. *Corresponding author. Email: m-wasielewski@northwestern.edu (M.R.W.); mdkrzyaniak@northwestern.edu (M.D.K.); stefano.carretta@unipr.it (S.C.) †These authors contributed equally to this work to substrates. Achieving this understanding would make the design of chiral molecular building blocks to manipulate electron spin states pos- sible, which has potential for QIS applications. In particular, the occurrence of CISS at the molecular level has been proposed as an en- abling technology for quantum applications, e.g., solving key issues like single-spin read- out and high-temperature spin qubit initial- ization (6). In this work, we show direct evidence of CISS in isolated covalent donor–chiral bridge–acceptor (D-Bc-A) molecules in which selective photo- excitation of D to its lowest excited singlet state (1*D) is followed by two rapid, sequential electron- transfer events: 1*D-Bc-A → D(cid:129)+-Bc(cid:129)–-A → D(cid:129)+- Bc-A(cid:129)– (Fig. 1A). If formation of D(cid:129)+-Bc-A(cid:129)– occurs in ≲1 ns and the effect of chirality is neglected, the resulting entangled electron spin pair is prepared initially in a nearly pure singlet state, 1(D(cid:129)+-Bc-A(cid:129)–). These states are commonly re- ferred to as spin-correlated radical pairs (SCRPs) and have been studied in systems ranging from photosynthetic proteins (18–21) and related mod- el systems (22–25) to DNA hairpins (26–30). However, in all these cases, no consideration was given to the possible influence of chirality on SCRP spin dynamics. To demonstrate the occurrence of CISS, we have synthesized pairs of covalent D-Bc-A enantiomers, (R)-1-h9 (-d9) and (S)-1-h9 (-d9), where D is either nondeuterated (-h9) or fully deuterated (-d9) peri-xanthenoxanthene (PXX) (31), Bc is a pair of naphthalene-1,8-dicarboximides that are linked at their 4-positions to form an enantiomeric pair of axially chiral dimers (R)-NMI2 and (S)-NMI2 (32), and A is naphthalene- 1,8:4,5-bis(dicarboximide) (NDI) (supplemen- tary materials; figs. S1 and S2). The structures of (R)-1-h9 (-d9) and (S)-1-h9 (-d9) and the corre- sponding achiral reference molecules 2-h9 (-d9) are shown in Fig. 1B. The enantiomers were sepa- rated by HPLC with a chiral stationary phase (fig. S3), and their circular dichroism spectra are given in fig. S4. We have characterized the charge separation and recombination dynamics of these molecules with transient optical absorption (TA) spectroscopy, and the CISS effect on their spin dynamics with time-resolved electron paramag- netic resonance (EPR) spectroscopy using either continuous (TREPR) or pulsed microwave ra- diation (pulse-EPR). We found that CISS yields characteristic features in the TREPR spectra of the photo- generated PXX(cid:129)+- NMI2-NDI(cid:129)– SCRP in (R)- 1-h9 (-d9) and (S)-1-h9 (-d9), which are absent in achiral 2-h9 (-d9), when the direction of electron transfer is oriented orthogonal to the applied static magnetic field direction, in agreement with simulations. Conversely, the corresponding spectra of PXX(cid:129)+-NMI2-NDI(cid:129)– are practically identical when the field is paral- lel to the electron-transfer direction. Time-resolved EPR spectroscopy Samples of (R)-1-h9 (-d9), (S)-1-h9 (-d9), and 2-h9 (-d9) were each prepared in the nematic liquid crystal 4-cyano-4'-(n-pentyl)biphenyl (5CB), which was aligned in a magnetic field at 295 K, then rapidly frozen to 85 K, which aligns the long axes of these molecules along the magnetic field. The orientation of the mol- ecules aligned in frozen 5CB can then be ro- tated relative to the applied magnetic field direction. Because solid 5CB is an optically scattering medium, to assess the photo-driven charge-separation dynamics of these molecules at low temperature, we used both femtosecond and nanosecond TA spectroscopy, substitut- ing glassy butyronitrile for 5CB at 105 K. Tran- sient absorption spectra and kinetics are given in figs. S5 and S6. The data show that in each case, ultrafast two-step charge separation occurs in ≲ 0.2 ns to give PXX(cid:129)+-NDI(cid:129)–, which recombines to its ground state in time constant t = 46 to 66 ms, providing ample time for TREPR measurements. The presence of a ~0.35-T sta- tic magnetic field in the TREPR experiments does not affect the ultrafast electron-transfer reactions because the Zeeman interaction (~0.3 cm−1) at that field strength is much less than the adiabatic energy gaps (~20 to 80 cm−1) for these reactions (see table S1 and the sup- plementary materials for details). We used pulse-EPR techniques to assess the quality of the alignment of (R)-1-h9 (-d9), (S)-1-h9 (-d9), and 2-h9 (-d9) in 5CB by mea- suring the isotropic exchange (J) and dipolar (D) spin-spin interactions for their photogen- erated SCRPs, where D(q) = d(1 – 3cos2q) and d = 52.04 MHz (cid:1) nm3=r3 DA in the point-dipole approximation, which gives detailed distance and orientation information as defined by the Hamiltonian in eq. S3. If photogenera- tion of the SCRP is followed by a Hahn echo microwave pulse sequence (p/2 pulse – delay t – p pulse – delay t – spin echo) and t is scanned, coherent oscillations between the eigenstates Eckvahl et al., Science 382, 197–201 (2023) 13 October 2023 1 of 5 RES EARCH | R E S E A R C H A R T I C L E iand FBj of the SCRP Hamiltonian FAj i (see below) that are related to both J and D(q) modulate the spin-echo amplitude (21, 33–37). When this experiment is performed on spin- coherent SCRPs, the echo appears out of phase— i.e., in the detection channel in quadrature to the one in which it is expected—and is there- fore termed out-of-phase electron spin echo en- velope modulation (OOP-ESEEM) (21, 33–37). For large SCRP distances, rDA, J can be neg- lected and the OOP-ESEEM oscillation fre- quency is approximately 2d when q = 0°, and d when q = 90°. Thus, OOP-ESEEM can be used to measure SCRP distances for a given angle of the dipolar axis relative to the magnetic field (28, 36–38). The dipolar axis in SCRPs con- nects the centroids of the spin distributions of the two radicals. Figures S7 and S8 show the OOP-ESEEM data for (R)-1-h9 (-d9), (S)-1- h9 (-d9), and 2-h9 (-d9), assuming that their dipolar axes are aligned parallel or perpen- dicular to the magnetic field. Fitting the OOP-ESEEM data showed that the measured (cid:129)+-NDI(cid:129)– distances of (R)-1-h9, (S)-1-h9, PXX-h9 and 2-h9 were 2.48 ± 0.01, 2.48 ± 0.01, and 2.28 ± 0.01 nm, respectively, whereas the cor- (cid:129)+-NDI(cid:129)– distances of (R)-1- responding PXX-d9 d9, (S)-1-d9, and 2-d9 were 2.53 ± 0.01, 2.51 ± 0.01, and 2.29 ± 0.02 nm, respectively (table S2). These experimental distances are con- sistent with the center-to-center distances between PXX and NDI determined from density functional-theory calculations on (R)- 1-h9, (S)-1-h9, and 2-h9, where rDA = 2.59, 2.60, and 2.40 nm, respectively (fig. S9 and tables S3 to S5). The agreement between the experimental and calculated distances shows that the D(cid:129)+-Bc-A(cid:129)– SCRPs are well- aligned along the magnetic field direction in frozen 5CB. The TREPR spectra of aligned (R)-1-h9, (S)- 1-h9, and 2-h9 were obtained by photoexciting the samples with a 450-nm, 7-ns laser pulse and monitoring the magnetization with con- tinuous microwaves by using direct detection (supplementary materials). The spectra ob- tained 100 ns after the laser pulse are shown in Fig. 2. When the long axes of these mol- ecules are aligned parallel to the magnetic field direction (q = 0°), both enantiomers as well as the achiral reference molecule gave the same spectra (Fig. 2, A and C). Rotating the samples so that the long axes of (R)-1-h9, (S)-1-h9, and 2-h9 are aligned perpendicu- lar to the magnetic field direction (q = 90°) resulted in the appearance of outer wings in the spectra of chiral (R)-1-h9 and (S)-1-h9 (Fig. 2B). No such enhancement was observed for achiral 2-h9. As explained below, we posit that these new features result from the con- tribution of CISS to the formation of the SCRPs in (R)-1-h9 and (S)-1-h9. Deuteration of PXX(cid:129)+ narrows the overall linewidth of (R)-1-d9, (S)-1-d9, and 2-d9 while retaining the same Fig. 1. Electron transfer pathways and molecular structures. (A) Electron transfer and intersystem-crossing pathways in a D-Bc-A system with no applied magnetic field, where kCS1 and kCS2 are the charge-separation rate constants, kST is singlet-triplet mixing rate constant, and kCRS and kCRT are the charge recombination rates through the singlet and triplet channels, respectively. (B) Structures of chiral (R)-1 and (S)-1 and achiral 2. The steric constraints imposed by linking the two NMI groups in (R)-1 and (S)-1 result in stable enantiomers that have axial chirality. orientation dependence of the signal (Fig. 2, C and D). Effect of CISS on radical pair spin dynamics In the molecules described here, the D(cid:129)+-A(cid:129)– distances are ≳23 Å, so that the spin-spin interactions J and D are small relative to the ~0.35-T applied magnetic field. Thus, the Zeeman term is by far the leading term in the SCRP spin Hamiltonian (eq. S3), so that the SCRP wavefunctions Sj i ¼ 1ffiffi ↑↓j Þ i p 2 Þ, which are magnetic ↑↓j and T0j field invariant, remain close in energy, where- i are well sepa- j i and T(cid:3)1 j as Tþ1 rated in energy from both Sj i and T0j i. In i are eigen- particular, both Tþ1 states of the spin Hamiltonian, whereas Sj i and T0j i are not eigenstates because of the different electronic g factors and hyperfine i and T(cid:3)1 i ¼ 1ffiffi p 2 i ¼ ↓↓j i ¼ ↑↑j i þ ↓↑j i (cid:3) ↓↑j i ð ð j j i yields FAj i ¼ cosf Sj i þ sinf T0j i ¼ (cid:3)sinf Sj i þ cosf T0j fields of the two spins. Coherent mixing of Sj i and T0j i and i (Fig. 3A), which FBj are eigenstates of the spin Hamiltonian, where the angle f in the mixing coefficients is de- rived from the magnetic parameters of the SCRP (supplementary materials) (39–41). i and FBj In the ultrafast electron-transfer regime ob- served here, the initial spin state for an achiral SCRP is the entangled singlet Sj i state that yields populations only on FAj i . Therefore, four allowed transitions occur be- i and the initially un- i and FBj tween FAj i states, giving rise j i and T(cid:3)1 j populated Tþ1 to a spin-polarized (out-of-equilibrium) EPR spectrum. When q = 0°, this results in a typ- ical (e, a, e, a) spin-polarization pattern (low to high field, where a = enhanced absorption and e = emission) because D(q) < 0 (39–41). Eckvahl et al., Science 382, 197–201 (2023) 13 October 2023 2 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 2. TREPR spectra. TREPR spectra of (R)-1-h9, (S)-1-h9, and 2-h9 (A and B) and (R)-1-d9, (S)-1-d9, and 2-d9 (C and D) oriented in the nematic liquid crystal 5CB at 85 K and 100 ns after a 450-nm, 7-ns laser pulse with the long axis of each molecule 0° [(A) and (C)] and 90° [(B) and (D)] relative to the applied magnetic field direction. (Insets) The spectra are shown with their intensities expanded to highlight features characteristic of CISS. TDAF, time after laser pulse. Conversely, when q = 90°, the pattern is re- versed (a, e, a, e) because D(q) > 0. Because the g tensors of PXX(cid:129)+ and NDI(cid:129)– are very similar— i.e., [2.0045, 2.0045, 2.0031] (42) and [2.0047, 2.0047, 2.0027] (43), respectively—the expected SCRP polarization patterns (e, a, e, a) or (a, e, a, e) are reduced to broadened (e, a) or (a, e) patterns, as observed experimentally for the achiral reference molecules 2-h9 and 2-d9 (Fig. 2, blue traces). Our OOP-ESEEM results show that the dipolar axis of each SCRP is well aligned with the long axis of each molecule so that the dipolar axis and the chirality axis of (R)-1-h9 (-d9) and (S)-1-h9 (-d9) are nearly parallel. The angle q between this axis and the applied magnetic field (B0) direction is depicted in Fig. 3, B and E, for the parallel and perpendicular orientations, respectively. CISS mixes triplet character into the initial singlet SCRP, thus the initial populations of FAj i and the corre- i , and T(cid:3)1 sponding transition intensities are predicted to change as well (44–47). If CISS is the sole contribution to the spin dynamics and q = 0° (Fig. 3D), then the state following electron i, depending on transfer would be ↑↓j the chirality of the enantiomer and whether B0 is parallel or antiparallel to the electron mo- tion. Given that the typical alignment of linear D-B-A molecules within nematic liquid crys- tals is not unidirectional, B0 has equal prob- ability of being parallel or antiparallel to the i or ↓↑j j i, Tþ1 i, FBj j electron motion and hence, if coherences are lost, the initial state is an equal mixture of ↑↓j i and ↓↑j i and is thus equivalent to having a pure initial Sj i state. This situation is shown schematically in Fig. 3C, where the blue and red traces depict the idealized TREPR spectra expected when CISS contributes 0 and 100%, respectively. Indeed, the observed spectra of both enantiomers as well as the achiral ref- erence molecule are practically identical for q = 0° (Fig. 2, A and C). j j i, and T(cid:3)1 By contrast, when the chirality axis is ortho- gonal to B0 (Fig. 3E), the initial state is very dif- ferent in the presence or absence of CISS. In particular, the CISS contribution initially pop- i (Fig. 3G and eq. S12). ulates Tþ1 Therefore, if the SCRP spin state has a 100% CISS contribution, the TREPR spectra have a nearly opposite intensity pattern with re- spect to the case in which CISS does not contribute. This is shown in Fig. 3F where the blue and red lines in the idealized TREPR spectra correspond to the intensity for the pure Sj i (IS) and pure CISS (ICISS) initial con- ditions, respectively. Starting from recent theoretical models (44–47) that describe the influence of CISS on SCRP spin dynamics in cases for which the CISS contribution is not 100%, the initial state will be a superposition or a mixture of Sj i, and T0j i along the chiral axis direction, making the detection of CISS less obvious. In fact, the spectral line intensities in this case are the weighted sum of IS and ICISS, which occur at the same resonance fields and tend to cancel out (see details in the supplementary materials). The key to unraveling CISS and pure singlet contributions to the SCRP spin state in the molecules studied here is the ob- servation of a larger EPR linewidth that occurs when CISS contributes. Indeed, the sum of the two contributions (Fig. 3F, black trace) yields a signal that displays lateral wings of opposite sign and central features that are narrower than those produced in the absence of CISS, exactly as observed experimentally in Fig. 2, B and D. These features are unambiguous signa- tures of CISS because they cannot be produced starting from an initial Sj i state, where the polarization pattern is fixed to (a, e) for q = 90° by the sign of D(q) (39–41). i and FBj The larger linewidth obtained for the CISS initial state arises from the very different dependence of IS and ICISS on the degree of coherent mixing in the eigenstates. Indeed, |ICISS| increases with increasing entanglement (f→0) whereas |IS| decreases (fig. S10). Explor- ing the variation of the intensity by varying the composition of the FAj i eigen- states is made possible by the presence of sev- eral nuclear spins and by distributed magnetic parameters, e.g., dipolar couplings, often termed strain. Therefore, moving from the center of each transition, i.e., the center of the distribu- tions of the magnetic parameters and hyperfine fields, to the tail of the lineshape corresponds to changing the composition of the eigenstates, which produces different linewidths for different initial states. If entanglement in the eigenstates is larger in the tails of the spectrum, the CISS contributions result in magnetic field–dependent broadening, giving rise to lateral contributions to the lineshape of opposite sign with respect to the central features (Fig. 3F, black trace). To confirm this interpretation, we considered the spectra of partially deuterated (R)-1-d9 and (S)-1-d9. By strongly diminishing the hyperfine couplings on one of the two radicals, we changed the distribution of the eigenstate composition and probed its effect on the lineshape. The mea- sured spectra for q = 0 and 90° are shown in Fig. 2, C and D, respectively. Although no quali- tative effect is visible in the parallel direction as expected, the lateral wings are substantially reduced in the perpendicular orientation. These spectra were simulated with a minimal SCRP model with either one spin-½ nucleus (hydro- gen atom) on both NDI(cid:129)– and PXX+ or only on NDI(cid:129)–, the latter of which is the partially deu- terated case. For reasonable values of the hyper- fine couplings, the simulations shown in Fig. 4 reproduce the experimental behavior. The intensities of the lateral wings are cor- rectly reproduced by combining IS and ICISS with weights of 41 and 59%, respectively (Fig. 4). Although a 59% CISS contribution is remarkable, Eckvahl et al., Science 382, 197–201 (2023) 13 October 2023 3 of 5 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. The CISS effect on the spin states of SCRPs. (A) SCRP spin states in the absence of CISS and in the presence of a static magnetic field B0 that is much greater than J, D, and the hyperfine interactions in both radicals for a singlet precursor. The enhanced absorptive (a) and emissive (e) microwave-induced EPR transitions are indicated. (B) Schematic of chiral molecules aligned parallel to B0. (C) TREPR spectra with q = 0° expected for an achiral SCRP (blue trace) and for a chiral SCRP with an initial state having a 100% (red trace) or partial CISS contribution (black trace). (D) SCRP spin states for q = 0° where the initial state has a 100% CISS contribution. (E) Schematic of chiral molecules aligned perpendicular to B0. (F) TREPR spectra with q = 90° expected for an achiral SCRP (blue trace) and for a chiral SCRP with an initial state having 100% CISS contribution (red trace) or a partial CISS contribution (black trace, rescaled). (G) SCRP spin states for q = 90° where the initial state has a 100% CISS contribution. The widths of the energy levels in (A), (D), and (G) indicate the population of the initial state, whereas the relative arrow thicknesses in the boxes depict the transition probabilities. it must be stressed that this is a minimal model in which the effect of the nuclei is accounted for only qualitatively, and a full spectral simulation with all nuclear spins in the fully protonated molecules is very demanding. However, we have been able to perform the simulation for the deuterated case, which includes all four 1H and two 14N nuclei coupled to the electron spin in NDI(cid:129)– and effects of dipolar strain. In this case, the experimental behavior is very well reproduced with a 47% CISS contribution, which is still considerable. Further evidence for the validity of this in- terpretation was obtained by investigating the time dependence of the TREPR spectra, which reflected the time evolution of the D(cid:129)+-Bc-A(cid:129)– spin states under the combined effect of coherent and incoherent terms as described by the stochastic Liouville equation and the presence of the mi- crowave field (supplementary materials). Indeed, figs. S11 and S12 show that the dependence of the observed intensity of the wings of the spectra are similar to that of the main peaks, which is in agreement with our numerical simulations. The CISS contribution to the spin dynamics of (R)-1-h9 (-d9) and (S)-1-h9 (-d9) is similar to the ~50% spin polarization that was recently reported for an axially chiral binaphthalene derivative covalently linked to a gold film de- posited on nickel (48). Although this single comparison suggests that the observed CISS effect for the binaphthalene attached to the gold surface may be largely due to the chiral molecule, Fig. 4. Simulations of the TREPR spectra with a minimal model of the SCRP. The model places one hydrogen nuclear spin-½ on both PXX(cid:129)+ and NDI(cid:129)– (A and B), or only on NDI(cid:129)– (C and D). The nuclear spins are coupled to each radical with isotropic hyperfine couplings aNDI = 6.3 MHz and aPXX = 10 MHz. (Insets) The simulations are shown with their intensities expanded to highlight features characteristic of CISS. The complete list of simulation parameters is given in table S7. Eckvahl et al., Science 382, 197–201 (2023) 13 October 2023 4 of 5 RES EARCH | R E S E A R C H A R T I C L E additional comparative work is needed on a variety of systems to warrant such a conclusion. Conclusions We have found direct evidence of the CISS effect on the spin dynamics of photogenerated radical ion pairs in molecular electron donor– acceptor molecules. The observation of CISS in these systems affords possibilities both for increasing our understanding of this important phenomenon and for its possible applications. These results show that the substrates or elec- trodes with their possibly large spin-orbit coupl- ings are not needed for CISS to occur, and that TREPR spectroscopy can directly access the spin dynamics that result from CISS. This provides key information to guide theoretical investigations and makes possible many new targeted experimental studies. In addition, observing CISS at the mo- lecular level is the first step required to trans- form this fundamental phenomenon into an enabling technology for quantum applications. RE FE RENCES AND N OT ES 1. M. Atzori, R. Sessoli, J. Am. Chem. Soc. 141, 11339–11352 (2019). 2. M. R. Wasielewski et al., Nat. Rev. Chem. 4, 490–504 (2020). 3. K. Ray, S. P. Ananthavel, D. H. Waldeck, R. Naaman, Science 283, 814–816 (1999). 4. R. Naaman, Y. Paltiel, D. H. Waldeck, Nat. Rev. Chem. 3, 250–260 (2019). 5. C. D. Aiello et al., ACS Nano 16, 4989–5035 (2022). 6. A. Chiesa et al., Adv. Mater. 35, e2300472 (2023). 7. S. H. Yang, R. Naaman, Y. Paltiel, S. S. P. Parkin, Nat. Rev. Phys. 3, 328–343 (2021). 16. C.-Z. Wang, V. Mujica, Y.-C. Lai, Nano Lett. 21, 10423–10430 (2021). 17. M. S. Zöllner, S. Varela, E. Medina, V. Mujica, C. Herrmann, J. Chem. Theory Comput. 16, 2914–2929 (2020). 18. R. Bittl, G. Kothe, Chem. Phys. Lett. 177, 547–553 (1991). 19. G. Kothe et al., Chem. Phys. Lett. 186, 474–480 (1991). 20. R. Bittl, A. van der Est, A. Kamlowski, W. Lubitz, D. Stehlik, Chem. Phys. Lett. 226, 349–358 (1994). 21. S. A. Dzuba, P. Gast, A. J. Hoff, Chem. Phys. Lett. 236, 595–602 (1995). 22. M. R. Wasielewski, G. L. Gaines III, G. P. Wiederrecht, W. A. Svec, M. P. Niemczyk, J. Am. Chem. Soc. 115, 10442–10443 (1993). 23. K. Hasharoni et al., J. Am. Chem. Soc. 117, 8055–8056 (1995). 24. D. Carbonera et al., J. Am. Chem. Soc. 120, 4398–4405 (1998). 25. S. M. Harvey, M. R. Wasielewski, J. Am. Chem. Soc. 143, 15508–15529 (2021). 26. R. Carmieli et al., J. Phys. Chem. A 113, 4691–4700 (2009). 27. R. Carmieli et al., J. Am. Chem. Soc. 134, 11251–11260 (2012). 28. J. H. Olshansky, M. D. Krzyaniak, R. M. Young, M. R. Wasielewski, J. Am. Chem. Soc. 141, 2152–2160 (2019). 29. J. H. Olshansky, J. Zhang, M. D. Krzyaniak, E. R. Lorenzo, M. R. Wasielewski, J. Am. Chem. Soc. 142, 3346–3350 (2020). 30. E. R. Lorenzo et al., J. Am. Chem. Soc. 143, 4625–4632 (2021). 31. J. A. Christensen, J. Zhang, J. Zhou, J. N. Nelson, M. R. Wasielewski, J. Phys. Chem. C 122, 23364–23370 (2018). 32. M.-T. Chen et al., Org. Chem. Front. 6, 3731–3740 (2019). 33. M. C. Thurnauer, J. R. Norris, Chem. Phys. Lett. 76, 557–561 (1980). 34. K. M. Salikhov, Y. E. Kandrashkin, A. K. Salikhov, Appl. Magn. Reson. 3, 199–216 (1992). 35. J. Tan, M. C. Thurnauer, J. R. Norris, Chem. Phys. Lett. 219, 283–290 (1994). 36. R. Bittl, S. G. Zech, J. Phys. Chem. B 101, 1429–1436 (1997). 37. S. Santabarbara et al., Biochemistry 45, 7389–7403 (2006). 38. R. Carmieli et al., J. Am. Chem. Soc. 131, 8372–8373 (2009). 39. C. D. Buckley, D. A. Hunter, P. J. Hore, K. A. McLauchlan, Chem. Phys. Lett. 135, 307–312 (1987). 40. G. L. Closs, M. D. E. Forbes, J. R. Norris, J. Phys. Chem. 91, 3592–3599 (1987). 41. A. J. Hoff et al., Spectrochim. Acta A Mol. Biomol. Spectrosc. 54, 2283–2293 (1998). 42. H. Mao, G. J. Pažėra, R. M. Young, M. D. Krzyaniak, M. R. Wasielewski, J. Am. Chem. Soc. 145, 6585–6593 (2023). AC KNOWLED GME NTS Funding: This work was supported by the National Science Foundation award CHE-2154627 (M.R.W.; synthesis, transient optical, and EPR measurements); the Center for Molecular Quantum Transduction, an Energy Frontier Research Center funded by the US Department of Energy, Office of Science, Basic Energy Sciences (BES) award DE-SC0021314 (M.D.K., EPR data analysis); and cofunded by the European Union (ERC-SyG CASTLE, project no. 101071533) (S.C.; calculations). However, views and opinions expressed are those of the authors only and do not necessarily reflect those of the European Union or the European Research Council. Neither the European Union nor the granting authority can be held responsible for them. Additional funding was provided by the Foundazione Cariparma (S.C.; calculations). 1H nuclear magnetic resonance spectroscopy and mass spectrometry were conducted in IMSERC facilities at Northwestern University, which have received support from the Soft and Hybrid Nanotechnology Experimental (SHyNE) Resource (NSF ECCS-2025633), NSF CHE-1048773, Northwestern University, the State of Illinois, and the International Institute for Nanotechnology. Author contributions: Conceptualization: M.R.W.; Methodology: H.J.E., N.A.T., J.M.B., M.D.K., A.C., S.C., R.M.Y., and M.R.W.; Investigation: H.J.E., N.A.T., J.M.B., M.D.K., A.C., S.C., R.M.Y., and M.R.W.; Visualization: H.J.E., N.A.T., A.C., R.M.Y., and M.R.W.; Funding acquisition: M.R.W. and S.C.; Project administration: M.R.W.; Supervision: M.R.W., M.D.K., and S.C.; Writing – original draft: M.R.W., M.D.K., H.J.E., N.A.T., S.C., and A.C.; Writing – review and editing: M.R.W., M.D.K., H.J.E., N.A.T., S.C., and A.C. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data are available in the main text, the supplementary materials, and Dryad (49). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www. science.org/about/science-licenses-journal-article-reuse. This research was funded in whole or in part by ERC-Synergy project CASTLE 101071533. The author will make the Author Accepted Manuscript (AAM) version available under a CC BY public copyright license. 8. R. Naaman, Y. Paltiel, D. H. Waldeck, Acc. Chem. Res. 53, 43. Y. Wu, M. D. Krzyaniak, J. F. Stoddart, M. R. Wasielewski, 9. 2659–2667 (2020). I. Carmeli, K. Senthil Kumar, O. Heifler, C. Carmeli, R. Naaman, Angew. Chem. Int. Ed. 53, 8953–8958 (2014). 10. J. Fransson, J. Phys. Chem. Lett. 13, 808–814 (2022). 11. J. Fransson, Nano Lett. 21, 3026–3032 (2021). 12. J. Fransson, Phys. Rev. B 102, 235416 (2020). 13. J. Fransson, J. Phys. Chem. Lett. 10, 7126–7132 (2019). 14. S. Naskar, V. Mujica, C. Herrmann, J. Phys. Chem. Lett. 14, 694–701 (2023). 15. A. Dianat et al., Nano Lett. 20, 7077–7086 (2020). J. Am. Chem. Soc. 139, 2948–2951 (2017). 44. T. P. Fay, D. T. Limmer, Nano Lett. 21, 6696–6702 (2021). 45. T. P. Fay, J. Phys. Chem. Lett. 12, 1407–1412 (2021). 46. J. Luo, P. J. Hore, New J. Phys. 23, 043032 (2021). 47. A. Chiesa et al., J. Phys. Chem. Lett. 12, 6341–6347 (2021). 48. D. Amsallem, A. Kumar, R. Naaman, O. Gidron, Chirality 35, 562–568 (2023). 49. H. J. Eckvahl et al., Direct observation of chirality-induced SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adj5328 Materials and Methods Supplementary Text Figs. S1 to S12 Tables S1 to S7 References (50–60) spin selectivity in electron donor-acceptor molecules, Dryad (2023); https://doi.org/10.5061/dryad.fbg79cp1r. Submitted 1 July 2023; accepted 23 August 2023 10.1126/science.adj5328 Eckvahl et al., Science 382, 197–201 (2023) 13 October 2023 5 of 5
10.1126_science.adj6491
RES EARCH INORGANIC CHEMISTRY An all-metal fullerene: [K@Au12Sb20]5− Yu-He Xu1, Wen-Juan Tian2, Alvaro Muñoz-Castro3, Gernot Frenking4,5, Zhong-Ming Sun1* The C60 fullerene molecule has attracted tremendous interest for its distinctive nearly spherical structure. By contrast, all-metal counterparts have been elusive: Fullerene-like clusters composed of noncarbon elements typically suffer from instability, resulting in more compact geometries that require multiple embedded atoms or external ligands for stabilization. In this work, we present the synthesis of an all-metal fullerene cluster, [K@Au12Sb20]5−, using a wet-chemistry method. The cluster's structure was determined by single crystal x-ray diffraction, which revealed a fullerene framework consisting of 20 antimony atoms. Theoretical calculations further indicate that this distinct cluster exhibits aromatic behavior. T he study of all-metal clusters reveals the delicate balance between electronic shells and structural geometry: The quantum confinement of electrons (1, 2) gives rise to a rich diversity of atomic arrange- ments with intriguing bonding characteristics (3–5). The discovery of buckminsterfullerene (C60), which marked a major milestone in the exploration and application of stable three- dimensional cages, has sprouted new research disciplines in chemical, physical, and material science (6, 7). The distinct near-spherical struc- ture of fullerenes along with the surface of delocalized p electrons produces many nota- ble properties and enables a wide range of applications in biology, medicine, electron- ics, and photovoltaics (8, 9). What’s more, the internal cavity of fullerenes provides a space for hosting a variety of atoms and molecules, giving rise to a class of endohedral clusters termed endofullerenes (10–13). The rapid prog- ress in fullerene-related clusters and extensive applications of fullerene-based materials have prompted the exploration of analogous hollow sphere molecules composed of other main- group or transition metal elements known as inorganic fullerenes (14). In74 with D3h sym- metry and In48Na12 with D3d symmetry are fullerene-like constructs found in the solid- state Zintl phase Na96In97Z2 (Z = Ni, Pd, Pt); both constitute the outermost shell of a four- layer onion-like structure rather than existing as hollow cages (15). Additionally, theoretical calculations have predicted the stability of an all-gold fullerene Au32, structurally very sim- ilar to C60 (16). However, experiments produced 8+ cluster featuring a compact configura- an Au32 tion of Au12@Au20, which was different from the previously anticipated fullerene, and the cationic cluster was protected by organic li- gand (17, 18). Another ligand-protected dodeca- hedral silafullerane was also reported, which encapsulated one chloride ion (19). Obtaining ligand-free C60 analogs with heavier atoms may be constrained by their susceptibility to re- arrangement into alternative, more stable structures, as evidenced by previous theo- retical studies. In this work, we report the isolation and characterization of an all-metal endohedral fullerene, [K@Au12Sb20]5− through a solution-based method in which only one K+ ion resides in a bare dodecahedral cage compris- ing 12 Au and 20 Sb atoms with distinct struc- tural features. Each Au atom sits in the center of an Sb pentagonal plane without breaking the structure of Sb20 cage, but rather stretching the cage size. The [K@Au12Sb20]5− cluster is held together exclusively by Au–Sb bonds exploit- ing the icosahedron-dodecahedron duality, thereby retaining an icosahedral, near-spherical geometry with similar size to C60, but com- posed of 32 atoms. Synthesis and characterization Compound [K(2,2,2-crypt)]5[K@Au12Sb20] was synthesized by reacting the Zintl phase K8SnSb4 with precursor Au(PPh3)Me in an ethylene- diamine solution, which was facilitated by the presence of [2.2.2]crypt (see the supplemen- tary materials). After stirring at room tem- perature for 7 hours, the color of the reaction 1State Key Laboratory of Elemento-Organic Chemistry, Tianjin Key Lab of Rare Earth Materials and Applications, School of Materials Science and Engineering, Nankai University, Tianjin 300350, China. 2Institute of Molecular Science, Shanxi University, Taiyuan 030006, China. 3Facultad de Ingeniería, Arquitectura y Diseño, Universidad San Sebastián, Bellavista 7, Santiago 8420524, Chile. 4Institute of Advanced Synthesis, School of Chemistry and Molecular Engineering, State Key Laboratory of Materials- Oriented Chemical Engineering, Nanjing Tech University, Nanjing 211816, China. 5Fachbereich Chemie, Philipps- Universität Marburg, Hans-Meerwein-Strasse 4, 35043 Marburg, Germany. *Corresponding author. Email: sunlab@nankai.edu.cn Fig. 1. Molecular structure of the [K@Au12Sb20]5– cluster. (A) Thermal ellipsoid plot (50% probability) of the cluster. (B) Front side view of (A). (C) Space-filling representation of the crystal structure. (D) Top view of (A). (E) A typical Sb5 pentagon face centered by a gold atom in the cluster (with Sb–Sb bond lengths marked). K is represented by cyan, Au by gold, and Sb by blue. Xu et al., Science 382, 840–843 (2023) 17 November 2023 1 of 4 RES EARCH | R E S E A R C H A R T I C L E mixture was observed to change from brown- red to brown-gold, and the stirring was stopped. The reaction vial was stored in a refrigerator at 10°C for about 5 days, and black block-shaped crystals were isolated from the bottom of the vial in a yield of 25% based on Au(PPh3)Me. The resulting complex was characterized by single- crystal x-ray diffraction, which revealed its crys- tallization in the triclinic space group P-1. The asymmetric unit contained five [K(2,2,2-crypt)]+ charge-balancing cations. Additionally, energy- dispersive x-ray spectroscopy (EDS) was used to determine the elemental composition of the compound. The obtained atom values for Au and Sb were in good agreement with the the- oretical values calculated for Au12Sb20. Many attempts were made to obtain mass spectral in- formation, but to no avail. Because of the high negative charge of the cluster, it is extremely unstable in the air, and a distinct gray precipi- tate was observed in the crystal solution once exposed to air, which generated great difficul- ties in mass spectral detection. As shown in Fig. 1, the anion [K@Au12Sb20]5− exhibits the overall structure of a slightly flat- tened dodecahedron with each Sb5 pentagonal face centered by one Au atom. The average short axis of the cage measures 7.30 Å (the distance between two opposing Au@Sb5 pen- tagonal faces) (Fig. 1, A to C), slightly exceeding the diameter of C60 (7.1 Å) (20), whereas the longest axis measures 9.03 Å (the distance be- tween the two farthest Sb atoms on the cage) (Fig. 1D), which is comparable to the calcu- lated diameter of the theoretically predicted Au32 (9.0 Å) (16). In all twelve planes, the sum of the angles between the central Au atom and the five vertices of the Sb5 ring is close to 360°, indicating that the gold atom lies on the sur- face of cyclo-Sb5 (in the Sb5 plane in Fig. 1E, for example, the sum of the angles is 359.9°). Such a planar pentacoordinate motif is unusual. Theoretical studies predicted a planar hexa- 2−, coordinate carbon atom in the anion CB6 which has not yet been synthesized (21). Sim- ilarly, an iron-centered planar cation [FeSb5]+ was predicted, in which the Sb5 ring is aroma- tic with equal-length Sb–Sb bonds of 2.973 Å (22). However, the Sb–Sb bond distances in [K@Au12Sb20]5− span a wide range from 3.114 to 3.436 Å (average of 3.227 Å), which are sig- nificantly longer than that of [FeSb5]+ as well as typical Sb–Sb single bonds (2.81 to 2.98 Å) (23–25), denoting distinct structural features. Additionally, the Sb20 dodecahedral shell is expanded compared with that of the reported [Sb@Ni12@Sb20]− compounds (where the aver- age Sb–Sb distance is 3.11 Å), potentially owing to the influence of the Au atom in the plane enlarging the Sb5 pentagon (26). All Au–Sb bonds are located on the faces of the dodeca- hedron, with a relatively narrow range of 2.698 to 2.798 Å, which is considerably longer than the bonds observed in [Au2Sb16]4− (2.67 to 2.71 Å) and [Sb3Au3Sb3]3− (2.59 to 2.61 Å) (27, 28). The central K+ ion is coordinated by twelve Au atoms, supporting their positioning on the Sb5 faces. The K–Au contacts range between 3.493 and 3.756 Å, indicating predominantly electrostatic interactions. Nevertheless, the presence of K+ as template cations remains crucial for the overall cluster stability. Theoretical analysis To gain insights into the chemical bonding in the [K@Au12Sb20]5− cluster, theoretical anal- ysis was conducted. The optimized structure for [K@Au12Sb20]5– revealed Au–Au distances of 4.002 Å, mediated by Au–Sb bonds of 2.773 Å, which compared well to the x-ray structure (Au–Au, 3.914 Ǻ; Au–Sb, 2.747 Ǻ). Comparison Fig. 2. AdNDP bonding patterns and canonical molecular orbitals for [K@Au12Sb20]5−. (A) Sb 5s, Au 5d lone pairs, and the four-center two-electron (4c–2e) s bonds over each Sb2Au2 quadrilateral. ON, occupation number. (B) Selected canonical molecular orbitals with their superatomic features (S, P, and D) indicated. Xu et al., Science 382, 840–843 (2023) 17 November 2023 2 of 4 RES EARCH | R E S E A R C H A R T I C L E Fig. 3. Magnetic behavior for [K@Au12Sb20]5–. (A) Three-dimensional and (B) contour-plot representation of NICS and the induced magnetic field under certain orientations of the external field. (C) Streamline representation from GIMIC calculations of the current density over spheres with radii of 3.0, 4.5, and 6.0 Å, with the 6.0-Å sphere given in side and top views, and a cut plane at the center of the spherical cluster that denotes the magnitude of current density vector field in nA/T. Isosurface values set at ±3 ppm. Blue represents shielding and red represents deshielding. between the calculated energies of a D5d-symmetry and Ih-symmetry structure indicated that the latter is favored by 5.4 kcal mol−1, suggesting that the experimentally characterized D5d- structure may be influenced by counterions and crystal packing effects. The resulting Au12 icosahedron enclosed an inner spherical cavity of diameter 7.491 Ǻ, which is significantly larger than the Au12 cage found in ligand- protected gold clusters of about 5.4 Å (29, 30). This suggests that the cage structure is sup- ported by Au–Sb bonds, providing a larger interior volume, which thus presents a promis- ing strategy for designing larger hollow clusters. The endohedral K+ atom was stabilized by a calculated encapsulation energy of –375.8 kcal mol−1, which was primarily driven by electro- static interactions that accounted for 90% of the stabilizing forces (table S6). The calculated highest occupied molecular orbital–lowest occupied molecular orbital (HOMO-LUMO) gap amounted to 2.57 eV at the hybrid PBE0 level. Vibrational analysis denoted a bouncing motion for the endohedral K+ atom between 70 and 30 cm−1. A theoretical comparison be- tween [K@Au12Sb20]5– and its hypothetical com- pact counterpart with Au–Au distances of 3.045 Å (fig. S7b) reveals an energetic preference for the characterized structure of 55.9 kcal mol−1. To effectively allocate the 238 valence elec- trons, we used the adaptive natural partition- ing (AdNDP) analysis with the AdNDP 2.0 code (31). The advantage of using the AdNDP method is its capacity to elucidate the chem- ical bonding arrangement, encompassing both Lewis bonding constituents [including lone pairs (1c2e) and two-center two-electron (2c2e) bonds] and delocalized bonding constituents. In addition to the 20 Sb 5s lone pairs and 60 Au 5d lone pairs, there are 30 pairs of 4c–2e s bonds distributed evenly over each butterfly- shaped Au2Sb2 quadrilateral, covering the sur- face of the [K@Au12Sb20]5– cluster (Fig. 2A). The occupation numbers of these bonds range from 1.91 to 1.94 |e|. The remaining 18 electrons are allocated to nine orbitals with superatomic features (S, P, and D), satisfying the 3D aromatic requirement of 2(n + 1)2 (n = 2) (Fig. 2B). These electron distributions contribute to the overall stability and distinct properties of the cluster. The natural atomic orbitals analysis provided valuable insights into the contribution of each atom to the orbitals (32). Table S4 presents the total contribution of each atom to these orbi- tals, revealing that nearly all atoms make sub- stantial contributions. The natural population analysis conducted on the optimized structure of [K@Au12Sb20]5– revealed that the central K atom carries a charge of +0.85 |e|, indicating the presence of electrostatic interactions between the inner K atom and the outer Au12Sb20 shell. Moreover, the detailed energy decomposition analysis results at the PBE0/STO-TZ2P-ZORA level, as shown in table S6, further support the ionic nature of the system. The analysis revealed that electro- static interactions dominate the K+-cage bond- ing, contributing more significantly (DEelstat, 89.6%) compared with orbital interactions (DEorb, 7.4%) in attracting local charges and stabiliz- ing the system. To evaluate the aromatic properties of [K@Au12Sb20]5–, the overall magnetic behavior was given by the three-dimensional represen- tation of nucleus-independent chemical shift (NICS) which accounted for the orientational- ly averaged behavior resulting from the exper- imental molecular tumbling in solution (Fig. 3). The NICS isosurface exhibits a shielding con- tour at the spherical cage, which suggests a spherical aromatic behavior (33, 34). To over- come the NICS exaltation near to heavy nuclei, we focused our analysis on the long-range char- acteristics of the induced magnetic field at the low-electron density limit. (35) Moreover, the representation of the magnetic response un- der specific orientation of the external field (Bind i; i = x, y, z) provides a picture of the shielding and deshielding regions that ac- count for the possible global aromatic char- acteristics in [K@Au12Sb20]5–. As a result, from Bind i, an enhanced long-range shielding region aligned to the applied field for different orien- tations was obtained, complemented with a de- shielding region in a perpendicular plane, which accounts for the shielding cone property inher- ent to aromatic species (29, 36). The long-range shielding region exhibited calculated values of –20.0 ppm at 7.5 Å from the center of the struc- ture, and of –2.6 ppm at 15.0 Å (Fig. 3b), thus supporting the spherical aromatic behavior of [K@Au12Sb20]5– and leading to an enhanced Xu et al., Science 382, 840–843 (2023) 17 November 2023 3 of 4 RES EARCH | R E S E A R C H A R T I C L E shielding region. In addition, the current density upon a z-aligned external field was given from gauge-including magnetically induced currents (GIMIC) calculations, denoting a collective of parallel currents around the cluster that were observed at inner regions (3.0 Å of radius), at the structure contour (4.5 Å), and outside of the spherical shell. This analysis supports the formation of a long-range shielding region ow- ing to the presence of aromatic currents upon an external field. Integration of the induced current strength denoted values of 9.8 nanoamperes per tesla (nA/T), contributed by +158.4 nA/T from diatropic and –148.6 nA/T from para- tropic currents, which is sizable in comparison to the prototypical planar aromatic species given by benzene, with a value of 12.1 nA/T at the PBE0/def2-tvpz level. At the PBE0/LanL2DZ level, a value of 21.2 nA/T was obtained, denot- ing dependence of the level of theory. The Au–Sb heterobonds play a crucial role in maintaining the structural integrity of the cage, whereas the endohedral cation acts as a template for structure formation. Future in- vestigations will focus on exploring alternative synthetic strategies that leverage the interplay between cage composition and endohedral templates, thereby enabling the rational and controlled synthesis of larger all-metal fullerenes. These superatoms hold great potential for the design and fabrication of precisely engineered nanostructures owing to their atomically pre- cise near-spherical structures. RE FE RENCES AND N OT ES 3. F. Lips, R. Clérac, S. Dehnen, Angew. Chem. Int. Ed. 50, 4. 960–964 (2011). J.-Q. Wang, S. Stegmaier, T. F. Fässler, Angew. Chem. Int. Ed. 48, 1998–2002 (2009). 5. M. Ruck, V. Dubenskyy, T. Söhnel, Angew. Chem. Int. Ed. 42, 2978–2982 (2003). 30. H. Yang et al., Nat. Commun. 4, 2422–2429 (2013). 31. D. Y. Zubarev, A. I. Boldyrev, Phys. Chem. Chem. Phys. 10, 5207–5217 (2008). 32. T. Lu, F. Chen, J. Comput. Chem. 33, 580–592 (2012). 33. N. V. Tkachenko, I. A. Popov, A. Muñoz-Castro, Eur. J. Inorg. Chem. 41, 4239–4250 (2021). 6. H. W. Kroto, J. R. Heath, S. C. Obrien, R. F. Curl, R. E. Smalley, 34. N. Fedik, A. I. Boldyrev, A. Muñoz-Castro, Phys. Chem. Nature 318, 162–163 (1985). Chem. Phys. 21, 25215–25219 (2019). 7. W. Kratschmer, L. D. Lamb, K. Fostiropoulos, D. R. Huffman, Nature 347, 354–358 (1990). 35. C. Foroutan-Nejad, Theor. Chem. Acc. 134, 8 (2015). 36. R. Islas, T. Heine, G. Merino, Acc. Chem. Res. 45, 215–228 8. K. Tanigaki et al., Nature 352, 222–223 (1991). 9. X. Lu, L. Feng, T. Akasaka, S. Nagase, Chem. Soc. Rev. 41, (2012). 7723–7760 (2012). 10. J. Zhao, Q. Du, S. Zhou, V. Kumar, Chem. Rev. 120, 9021–9163 (2020). 11. S. Aoyagi et al., Nat. Chem. 2, 678–683 (2010). 12. C. M. Beavers et al., J. Am. Chem. Soc. 133, 15338–15341 (2011). 13. S. Stevenson et al., Nature 401, 55–57 (1999). 14. M. J. Moses, J. C. Fettinger, B. W. Eichhorn, Science 300, 778–780 (2003). 15. S. C. Sevov, J. D. Corbett, Science 262, 880–883 (1993). 16. M. P. Johansson, D. Sundholm, J. Vaara, Angew. Chem. Int. Ed. 43, 2678–2681 (2004). 17. S. Kenzler et al., Angew. Chem. Int. Ed. 58, 5902–5905 (2019). 18. S.-F. Yuan, C.-Q. Xu, J. Li, Q.-M. Wang, Angew. Chem. Int. Ed. 58, 5906–5909 (2019). 19. J. Tillmann et al., Angew. Chem. Int. Ed. 54, 5429–5433 (2015). 20. W. I. F. David et al., Nature 353, 147–149 (1991). 21. K. Exner, P. von Ragué Schleyer, Science 290, 1937–1940 (2000). 22. M. Lein, J. Frunzke, G. Frenking, Angew. Chem. Int. Ed. 42, 1303–1306 (2003). 23. M. Westerhausen, S. Weinrich, P. Mayer, Z. Anorg. Allg. Chem. 629, 1153–1156 (2003). 24. H. Breunig, R. Rösler, Coord. Chem. Rev. 163, 33–53 (1997). 25. S. C. Critchlow, J. D. Corbett, Inorg. Chem. 23, 770–774 (1984). 26. Y. Wang et al., J. Am. Chem. Soc. 139, 619–622 (2017). 27. I. A. Popov et al., Angew. Chem. Int. Ed. 55, 15344–15346 (2016). 28. F.-X. Pan et al., J. Am. Chem. Soc. 137, 10954–10957 AC KNOWLED GME NTS We are grateful for the valuable discussions with X.-B. Wang (PNNL), J. Xu (NKU), and N. Li (Rigaku Beijing Co., Ltd.). Funding: This work was supported by the National Natural Science Foundation of China (nos. 92161102 and 22371140), the Natural Science Foundation of Tianjin City (no. 21JCZXJC00140), and 111 project (B18030) from China (MOE). A.M.-C. acknowledges financial support from ANID FONDECYT Regular 1221676. Author contributions: Conceptualization: Z.-M.S.; Methodology: A.M.-C. and Z.-M.S.; Investigation: Y.-H.X.; Visualization: Y.-H.X., W.-J.T., A.M.-C., and Z.-M.S.; Funding acquisition: A.M.-C. and Z.-M.S.; Project administration: Z.-M.S.; Supervision: Z.-M.S.; Writing – original draft: Y.-H.X, W.-J.T., A.M.-C., and Z.-M.S.; Writing – review and editing: Y.-H.X, W.-J.T., A.M.-C., G.F., and Z.-M.S. Competing interests: The authors declare no competing interests. Data and materials availability: X-ray data are available free of charge from the Cambridge Crystallographic Data Centre under reference numbers CCDC 2269174 (method 1) and 2291088 (method 2). All other experimental, spectroscopic, crystallographic, and computational data are included in the supplementary materials. License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science-licenses-journal- article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adj6491 Materials and Methods Figs. S1 to S9 Tables S1 to S6 References (37–53) 1. A. W. Castleman Jr., S. N. Khanna, J. Phys. Chem. C 113, (2015). 2664–2675 (2009). 29. X. Kang, H. Chong, M. Zhu, Nanoscale 10, 10758–10834 2. H. Häkkinen, Chem. Soc. Rev. 37, 1847–1859 (2008). (2018). Submitted 8 July 2023; resubmitted 3 September 2023 Accepted 4 October 2023 10.1126/science.adj6491 Xu et al., Science 382, 840–843 (2023) 17 November 2023 4 of 4
10.1126_science.adi3038
RES EARCH VOLCANOLOGY Fast and destructive density currents created by ocean-entering volcanic eruptions Michael A. Clare1*†, Isobel A. Yeo1*†, Sally Watson2, Richard Wysoczanski2, Sarah Seabrook2, Kevin Mackay2, James E. Hunt1, Emily Lane2, Peter J. Talling3, Edward Pope3, Shane Cronin4, Marta Ribó5, Taaniela Kula6, David Tappin7, Stuart Henrys8, Cornel de Ronde8, Morelia Urlaub9, Stefan Kutterolf9, Samuiela Fonua10, Semisi Panuve10, Dean Veverka11, Ronald Rapp12, Valey Kamalov13, Michael Williams2 Volcanic eruptions on land create hot and fast pyroclastic density currents, triggering tsunamis or surges that travel over water where they reach the ocean. However, no field study has documented what happens when large volumes of erupted volcanic material are instead delivered directly into the ocean. We show how the rapid emplacement of large volumes of erupted material onto steep submerged slopes triggered extremely fast (122 kilometers per hour) and long-runout (>100 kilometers) seafloor currents. These density currents were faster than those triggered by earthquakes, floods, or storms, and they broke seafloor cables, cutting off a nation from the rest of the world. The deep scours excavated by these currents are similar to those around many submerged volcanoes, providing evidence of large eruptions at other sites worldwide. E xplosive volcanism poses a wide range of hazards, with more than a third of vol- canic fatalities attributed to fast (up to hundreds of kilometers per hour) and high-temperature pyroclastic density cur- rents triggered by phreatic explosions, pyro- clastic fountaining, lateral blasts, caldera and dome collapses, and the vertical collapse of eruption columns where they contact the ground (1–6). Study of the behavior of pyro- clastic density currents on land has revealed a spectrum from dense to dilute and turbulent modes of flow, across which a range of hazards exists (1–3). This spectrum also relates to other types of particulate density currents, including snow avalanches and underwater sediment– laden flows called turbidity currents (7–9). Where terrestrially initiated pyroclastic density currents reach the ocean, they create different hazards. Such currents can generate tsunamis, create phreatic explosions as hot currents in- teract with water, travel over the sea, and/or rapidly cool and transition into a turbidity current, damaging seafloor infrastructure 1National Oceanography Centre, Southampton, UK. 2National Institute of Water and Atmospheric Research (NIWA), Auckland, Aotearoa New Zealand. 3Department of Geography and Department of Earth Sciences, Durham University, Durham, UK. 4School of Environment, University of Auckland, Auckland, Aotearoa New Zealand. 5Department of Environmental Science, Auckland University of Technology, Auckland, Aotearoa New Zealand. 6Ministry of Lands and Natural Resources, Nuku‘alofa, Kingdom of Tonga. 7British Geological Survey, Keyworth, UK. 8GNS Science, Lower Hutt, Aotearoa New Zealand. 9GEOMAR Helmholtz Centre for Ocean Research Kiel, Kiel, Germany. 10Tonga Cable Ltd, Nuku‘alofa, Kingdom of Tonga. 11Southern Cross Cable Network, North Ryde, New South Wales, Australia. 12SubCom, Newington, NH, USA. 13Valey Kamalov LLC, Gainesville, FL, USA. *Corresponding author. Email: michael.clare@noc.ac.uk (M.A.C.); i.yeo@noc.ac.uk (I.A.Y.) †These authors contributed equally to this work. and devastating marine biological communi- ties (10–15). Repeat terrestrial surveys and sampling after the occurence of pyroclastic density cur- rents have enabled reconstruction of flow prop- erties from the resultant deposits, characterizing the associated hazards (16–18). However, field- scale observations of density currents linked to volcanic eruptions in marine settings are rare, owing to often-remote locations and in- accessibility of in situ deposits. The behavior of terrestrially initiated pyroclastic density currents that cross land to enter the ocean has only been documented at a single field site after a small (0.19 km3) eruption (12–14), where- as no equivalent study has shown what happens when an eruption directly delivers volcanic material into the ocean. We address this knowl- edge gap with observations of underwater vol- caniclastic density currents triggered during the eruption of the partially submerged Hunga Tonga–Hunga Ha‘apai volcano (hereafter re- ferred to as Hunga volcano) in the Kingdom of Tonga. We use the term “volcaniclastic den- sity current,” which encompasses a spectrum of density currents linked to a volcanic erup- tion, from hot gas–supported pyroclastic den- sity currents to cold fluid–supported turbidity currents. A key control on the hazard posed by any type of density current is its velocity (9, 19–22). Although recent advances in technology have enabled the direct measurement of turbidity currents and snow avalanches, no velocity measurements exist for underwater volcani- clastic density currents (9, 19–21). These lim- itations are notable, considering the distinct hazards posed by partially or fully submerged volcanoes, which account for more than three- quarters of active volcanoes worldwide (6). Consequently, our knowledge relies on studies of ancient ocean-entering eruptions (23, 24), scaled-down laboratory modeling (25), and anal- ysis of geomorphic features around submerged volcanoes to infer the behavior of past erup- tions (26, 27). Fields of large sediment waves and scours, commonly observed radiating around submerged flanks of volcanoes, are thought to be diagnostic of catastrophic eruptions (26–28). However, this hypothesis remains untested be- cause of a lack of repeat seafloor surveys before and after a large eruption. These uncertain- ties severely limit the understanding of the behavior and associated risks at submerged volcanoes. We present observations of voluminous vol- caniclastic density currents that were triggered by the 15 January 2022 eruption of Hunga vol- cano in the Kingdom of Tonga. This eruption was the most explosive in more than a century and had worldwide impacts (29–35). The erup- tion plume entered the mesosphere (57 km high), tsunamis traveled across the Pacific Ocean and caused 19- to 20-m runups in Tonga, and a pressure wave encircled the globe mul- tiple times (29–31, 33, 34). More than 1 hour after the main eruption, Tonga’s only interna- tional subsea telecommunications cable was severed (Fig. 1), disconnecting the entire nation from global digital communications at a critical time for disaster response (36). Such an inci- dent has wider implications because subsea cables carry >99% of all international data traffic, including the internet and trillions of dollars per day in financial transactions (37). The >6 km3 eruption expelled a volume of material equivalent to the annual sediment flux from all the world’s rivers combined, much of which directly entered the ocean through eruption column collapses (15, 38). The rapid escalation in explosivity and the resultant hazards were unexpected, exposing a gap in understanding of many similar, yet unmoni- tored, volcanoes along the Tonga-Tofua arc and volcanic settings worldwide (6, 29, 30, 39). By integrating datasets that document their timing and extent, we determined the behavior of underwater volcaniclastic density currents triggered by the 2022 eruption of Hunga vol- cano. These currents traveled >100 km and caused extensive damage to seafloor cables, from which we could estimate their velocity, which reached up to 122 km/hour. These cur- rents differ markedly from those triggered by terrestrially initiated pyroclastic density currents that enter the ocean, and their velocity is higher than that recorded for other underwater den- sity currents, including those triggered by ter- restrial floods, large earthquakes, or ocean waves (Table 1). The currents created distinc- tive scours around Hunga volcano, excavating >100-m-deep regions into the volcanic edifice (Fig. 2), which were evident by comparing sur- veys 3 months after the eruption to pre-eruption surveys performed in 2016. Bedforms like these Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 1 of 7 RES EARCH | R E S E A R C H A R T I C L E are generated by, and thus probably record, large explosive eruptions. Extensive damage to seafloor cables At 03:47 (all times UTC) on 15 January 2022, a low eruptive plume was observed at Hunga volcano, marking the onset of the main erup- tive phase, with a steady narrow plume rising to a height of >10 km (34) (Fig. 1). At 04:15 (T0a), a large explosion occurred [volcanic ex- plosivity index of between 5 and 6 (35)], which transformed the plume. High eruption rates rapidly increased the height and width of an expanding umbrella-shaped eruption plume (34) (fig. S6). The plume reached a height of between 16 and 18 km at 04:17. A second major explosion (T0b) occurred at 04:21, with addi- tional explosions generating sonic booms until 04:25 (inception of the atmospheric pressure wave). By this time, the umbrella cloud had expanded to a width of at least 120 km, and the central plume was >15 km wide. Plume collapses into the ocean below the umbrella cloud occurred soon after 04:17, especially from 04:20 onward (34) (fig. S6). By 04:50 to 04:55, the central high plume ceased rising and was dispersed by the wind. However, the eruption continued vigorously with a lower plume (~17 to 21 km high) formed beneath the larger 57-km-high plume. Both subsea cables laid near Hunga volcano were broken on 15 January, but the timing of this damage lagged after the two most in- tense explosions (T0a and T0b) by 9 to 15 min for the domestic cable (at 04:30) and by 83 to 89 min for the international cable (05:44) (table S1). The timing of these breaks is known to the nearest minute, on the basis of loss of data transmission, ultimately with complete loss of internet capacity when the international cable was severed. The dis- tance of the first point of cable damage from shore was determined immediately using optical time-domain reflectometry, but the full extent could not be assessed until a cable repair ship retrieved the intact ends of the cable on either side of the damaged zone. The international cable repair took 5 weeks, as the closest repair ship was 2500 km away, in Papua New Guinea, and >100 km of replace- ment cable was required. Communications were limited across the kingdom until the domestic cable was finally repaired, 18 months after the eruption. Prior to repair, cable damage was thought to be caused by local seabed landslides; however, the extent of damage was far greater than antic- ipated. More than 89 km of the international cable was broken and/or buried to a depth beyond recovery, while 105 km of the domestic cable was affected. Moreover, the international cable was recovered at a distance of 5 km to the north of its originally laid position, closer toward the volcano. This incident was the largest length Fig. 1. Extensive damage caused to seafloor cables by volcaniclastic density currents generated by column collapse at Hunga volcano. (A) Locations and extent of cable damage on the domestic and international seafloor cables resulting from the volcaniclastic density currents (pathways shown as arrowed lines) plotted on bathymetric data acquired 3 months after the eruption. The thickness of volcaniclastic density current deposits as sampled by multicoring is depicted as size-scaled solid circles, showing that these currents deposited material at least 108 km away from the caldera. Actual sampled thickness of volcaniclastic density current deposit is annotated. Where the base of the density current deposit was not sampled, the thickness is given as >X cm. (B) Internet capacity shown for typical periods (in gray) compared with the sudden loss of internet traffic, which flatlined at 05:44 on 15 January 2022, when the international cable was broken. (C) Enhanced timeline of the main eruptive phase of Hunga volcano on 15 January 2022, including the two major eruptions that caused ocean-entering column collapses. The timings of the two cable breaks are marked with stars, showing that they occurred after the main explosive eruptions. Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 2 of 7 RES EARCH | R E S E A R C H A R T I C L E of cable damage since telegraph cables were buried by the >200-km3 Grand Banks landslide offshore Newfoundland in 1929 (40) and the longest single cable repair in the modern fiber- optic era (Fig. 1). We did not find evidence within the resolution of imaging of slope fail- ures on the deep-water volcano flanks around the cables and surrounding slopes. We therefore relate the cable damage to powerful and long runout volcaniclastic density currents triggered by the effective delivery of large volumes of erupted pyroclastic material directly into the ocean. Insights into underwater density currents triggered by eruption column collapse Dense and highly erosive proximal regime Linear gullies and trains of crescentic scours, incised up to 100 m, radiate from the caldera (Fig. 2), accounting for 3.5 km3 of erosion (15) and representing an additional removed mass of >50% of the original erupted volume. Ero- sional features include scours of up to 2 km in width and upslope-asymmetric bedforms (30 to 60 m wave height, 500 to 2000 m wavelength) seen in pre-eruption surveys, which migrated up to 1.5 km upslope. Pronounced erosion observed from the post-eruption survey occurs within a radius of <9.2 km from the caldera rim and only on the steepest slopes (>10°, locally >40°; Fig. 2). Recovery of material by seafloor coring was not successful on such proximal slopes, presumably owing to the presence of competent Table 1. Reported velocities of fast-moving (>4 m/s) submerged particulate density currents worldwide. Values based on sequential seafloor cable breaks or acoustic monitoring. Also compared are subaerial density currents, including pyroclastic density currents and snow avalanches. Location Volume Interpreted trigger Minimum runout distance (km) Maximum recorded transit velocity (m/s) Velocity calculation based on Submarine particulate density currents, including turbidity currents and volcaniclastic density currents ............................................................................................................................................................................................................................................................................................................................................ 1979 Nice Airport, Mediterranean (58) 0.008 km3 Mediterranean (59) Not known 1929 Grand Banks landslide, Newfoundland (40) >200 km3 Canyon, Taiwan (37) earthquake, Algeria (60) Canyon, Taiwan (37) Not known Not known Not known 2009 Gaoping Canyon, Taiwan (37) Not known ............................................................................................................................................................................................................................................................................................................................................ Gioia Canyon, 15 4.5 ............................................................................................................................................................................................................................................................................................................................................ 120 7 Seafloor cable breaks 800 19.1 Seafloor cable breaks ............................................................................................................................................................................................................................................................................................................................................ 2006 Gaoping ............................................................................................................................................................................................................................................................................................................................................ 1954 Orleansville ............................................................................................................................................................................................................................................................................................................................................ 2009 Gaoping ............................................................................................................................................................................................................................................................................................................................................ Not known 380 380 20 20.5 10.3 ............................................................................................................................................................................................................................................................................................................................................ 2020 Deep-sea Cable breaks and moored acoustic Doppler current profiler array ............................................................................................................................................................................................................................................................................................................................................ Low spring tide after large river flood Congo Canyon, West Africa (19) 2.675 km3 1130 8 650 16.6 Seafloor cable breaks 50 7.2 Moored acoustic Doppler current profiler array 50 8 2015–17 Monterey Canyon, California (55) Not known Canyon, northeast Atlantic (61) 2022 Hunga volcano, Tonga (this study) Not known >6.3 km3 [based on deposited volume on slopes outside of the caldera (15)] ............................................................................................................................................................................................................................................................................................................................................ 2019–20 Whittard ............................................................................................................................................................................................................................................................................................................................................ Eruption column collapse >100 33.8 Cable breaks ............................................................................................................................................................................................................................................................................................................................................ Subaerial particulate density currents ............................................................................................................................................................................................................................................................................................................................................ Largest snow avalanches (22, 62) 0.01 km3 Various 3 to 5 70 ............................................................................................................................................................................................................................................................................................................................................ Terrestrial pyroclastic density currents Dome or flank collapse or phreatomagmatic eruption Up to tens of kilometers 7 to 210 See table S4 (52) Up to 5.5 km3 for those with reported speeds, but can exceed hundreds of cubic kilometers ............................................................................................................................................................................................................................................................................................................................................ Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 3 of 7 Construction activity; slope failure during airport extension Construction activity; slope failure near port Magnitude 7.2 earthquake triggered continental slope failure Magnitude 7.0 Pingtung earthquake Magnitude 6.7 Orleansville earthquake Large river flood after typhoon Large river flood during typhoon Oceanographic trigger related to along-shelf transport Oceanographic trigger related to across-shelf transport Seafloor cable breaks Seafloor cable breaks Seafloor cable breaks Seafloor cable breaks Moored acoustic Doppler current profiler array Radar and pressure plate measurements RES EARCH | R E S E A R C H A R T I C L E B +100 ] m [ e g n a h c n o i t a v e l E 0 -100 B 0 E' A B' Depositional obel C' C 8 4 Distance [km] Distance [km] 0.5 1.0 ' C 12 Deeply incised gully B B B D' D' D' D D D Collapsed caldera F F F -60 D D' Erosional regime Depositional regime Domestic cable 20 10 0 F ' 6 Distance [km] 12 ] s e e r g e d [ t n e i d a r G Erosion Deposition Erosional regime Depositional regime 6 Distance [km] 12 ]20 s e e r g e d [ 10 10 0 0 t n e i d a r G ' B Depositional obe l E n o i t a v e l E E ] m [ e g n a h c Crescentic scours Up to 22 m deposition on domestic cable 20 E E' 0 0 Distance [km] 4 F' ] m [ e g n a h c n o i t a v e l E >+50 m <-50 m ' S " 0 0 4 ° 0 2 4 km 175°20'0"W H NE channel E channel NW channel International cable Other volcanoes (Pope et al., 2017) Hunga volcano (This Study) ] m [ t h g i e h e v a W 100 10 1 Small-scale bedforms Large-scale scours Large-scale bedforms - 001 0 100 10 100 1000 10000 Change in seabed elevation [m] Wavelength [m] Fig. 2. Sculpting of the seafloor by powerful volcaniclastic density cur- rents on the slopes proximal to Hunga volcano. (A) Elevation-difference map generated between pre- (2017) and post-eruption (April to June 2022) bathymetric surveys shows localized but major seafloor changes. The domestic cable is shown as a red line. (B to F) Elevation changes shown for selected locations in cross section, including the incision of deep (locally >120 m) gullies and upslope-migrating crescentic scours on steep (>10°) slopes [(B), (E), and (F)] and the deposition of thick (up to 40 m) lobes [(C) and (D)] where slopes shallow. (G) Cross plot of change in seabed elevation and local seafloor gradient (based on pre-eruption bathymetry) illustrating how erosion dominates on the steepest slopes, whereas deposition is largely restricted to slopes <10°. (H) Comparison of bedform morphometrics observed on the proximal flanks of Hunga volcano and around the area of the damage to the international cable with those from a database based on 17 volcanic islands worldwide (8) to show that the large-scale scours and bedforms plot within the range of those previously interpreted to relate to large explosive eruptions. Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 4 of 7 C n o i t a v e l E ] m [ e g n a h c 40 C 20 0 00 0 0 D n o i t a v e l E ] m [ e g n a h c -120 F ] m [ e g n a h c n o i t a v e l E +100 0 F -100 0 G ] s e e r g e d [ e p o l s d e b a e s l a c o L 30 20 10 0 -200 RES EARCH | R E S E A R C H A R T I C L E A 40 <18 km from Hunga volcano Runout not reported <70 km from Hunga volcano 30 20 10 0 1 ] s / m [ y t i c o l e v d e d r o c e r . x a M B Initiation Mechanism Earthquake River flood Oceanographic Construction Eruption column collapse 120 80 40 / ] r h m k [ y t i c o l e v d e d r o c e r . x a M 10 1000 100 Minimum runout distance [km] 0 10000 Eruption column collapse at submerged edifice e.g. Hunga volcano Rapid collapse of large volume and tall eruption column >6 km3 Up to 45 degree slope Direct, vertical entry of pyroclastic the sea material into up to 122 km/hr C River flood e.g. Congo Canyon, W Africa Gaoping Canyon, SW Taiwan <5 degree slope A) Plunging of dense sediment-laden flood water or B) Flushing event during low Spring tide after flood Up to ~3 km3 up to ~70 km/hr D Continental slope failure e.g. Grand Banks, Newfoundland Externally-triggered (e.g. earthquake or construction) landslide Up to 100s km3 up to ~70 km/hr <5 degree slope E Subaerial dome collapse e.g. Montserrat, Caribbean Subaerial pyroclastic density currents travel over land and then enter the sea <1 km3 <8 degree slope Velocity not known F Volcanic flank collapse e.g. Hawai’i, Canary Islands Failure of volcanic flank, generates a density flow as collapsed material disintegrates No reported velocities Up to 100s km3 10s of degree slope Fig. 3. Density currents triggered by the Hunga volcano eruption are the fastest reported for any submerged particulate density current to date. (A) Measured velocities and minimum runout distances for different underwater particulate density currents, categorized according to their triggers (as detailed in Table 1). Precise runout distances cannot be presented, as the current often ran out beyond the monitoring array or the location of seafloor cables. (B to F) Schematics illustrating the inception mechanisms for different submerged density currents: (B) rapid-eruption column collapse, causing vertical plunging of dense pyroclastic material onto an exceptionally steep slope, as seen at Hunga volcano; (C) river flood–triggered turbidity currents that enter the ocean either laterally (where sediment is flushed offshore) or obliquely (where dense sediment–laden flood water plunges), as observed in the Gaoping and Congo canyons; (D) continental slope collapses that are initiated by external ground disturbances, such as large earthquakes or construction activity, which can initiate on very low angle slopes; (E) initiated by subaerial volcanic dome collapse entering the ocean obliquely, as in the case of Montserrat; (F) initiated by the collapse of volcanic island flanks. volcanic rock and/or coarse granular material. Deposition occurs as slope angles reduce (<10°), where well-defined lobes (up to 40 m thick and 7 km wide) accumulated downstream of two of the erosional chutes. As this lobate deposition occurs on relatively steep slopes, a dense granular current with high basal fric- tion was likely responsible for the proximal depositional lobes (41–43). This is in stark contrast to turbidity currents, which typically do not deposit on steep slopes or, where they do, leave only very thin deposits (44). Indeed, some turbidity currents only deposit on slopes of less than 0.05° (45). The intense erosion on steep slopes (Fig. 2) caused currents to increase their sediment mass substantially, thereby enhancing their power and mobility (46). For context, the eroded volume is 0.5 km3 greater than the largest known his- torical volcanic landslide [3 km3, Mount St. Helens (47)] and ~1 km3 greater than the sedi- ment volume eroded by the longest monitored turbidity current that traveled >1000 km along the deep-sea Congo Canyon (19). Stepped trains of upslope-migrating crescentic scours and large- scale bedforms on steep slopes evidence Froude- supercritical currents undergoing a series of hydraulic jumps (48–50). Similar scours and bedforms are a common feature proximal to the initiation of other large-volume particulate density currents, including those in nonvolcanic settings [e.g., the 1 km3 earthquake-triggered turbidity current in Kaikōura Canyon (51)] and are of similar scale to those thought to diagnose past large explosive eruptions on submerged volcanic flanks (25–28) (Fig. 2H). We confirm this previously hypothesized link, showing that multiple radial chutes and bedforms can be formed during an individual explosive erup- tion, which has important implications for as- sessing hazards from seabed geomorphology at other submerged volcanoes worldwide. Depositional distal regime The remaining surveyed area is characterized by widespread, relatively featureless blanketed deposition, with an average of +2.8 m eleva- tion change (i.e., net depositional). In the valley southeast of Hunga volcano (location of the domestic cable), slopes rapidly reduce to <1.5°, and ponded deposition occurred (up to 27 m thick). Even where elevation change was not resolvable from repeat seafloor surveys in deep water, sampling (up to 108 km from Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 5 of 7 RES EARCH | R E S E A R C H A R T I C L E the caldera) recovered volcaniclastic deposits emplaced by density currents (Fig. 1). Density current deposits are generally of decimeter thickness; comprise sand to granule–sized vol- canic material that fines upward, with internal lamination, ripples, and sharp bases; and are geochemically linked to the 2022 Hunga vol- canic eruption (16). Those deposits are over- lain by thinner deposits interpreted as ash fallout (52) (fig. S2 and table S2). The area around the international cable was not cov- ered by preexisting bathymetric data; how- ever, detailed seafloor backscatter data reveal trains of bedforms (1 to 2 m wave height, 100 to 200 m wavelength) within a valley be- tween seamounts (fig. S1). These bedforms evidence a complex flow pathway, as corrob- orated by modeling (15). So intense was the topographic steering, that the cable was moved 5 km toward the volcano by the density current (fig. S1). Contrasting behavior of underwater volcaniclastic density currents These depositional and erosional patterns con- trast with observations from repeat seafloor surveys and sampling of terrestrially initiated pyroclastic density currents that entered the ocean offshore Montserrat in the Caribbean (12–14). Deposits offshore Montserrat formed in two parts: (i) coarse-grained ridges up to 60 m thick within 3 to 4 km of the ocean- entry point formed by a dense granular current; and (ii) a broader deep-water lobe, comprising centimeters-thick fine-grained deposits related to a dilute turbidity current (12–14). The prox- imal deposition of ridges offshore Montserrat contrasts with the erosional chutes, scours, and asymmetric bedforms we observe on Hunga volcano. Further, the lobes at Hunga volcano are much higher relief, thicker (tens of meters thick rather than centimeters thick), on steeper slopes (up to between 8° and 10° compared with <2°), and likely composed of coarser material, with deposits sampled at least 108 km from the edi- fice, compared with 40 km offshore Montserrat (Fig. 1). Seafloor cores show that coarse granular deposits were deposited at least 80 km from the Hunga edifice, indicating that currents maintained high densities over this distance, also in contrast with the dilute flow origin for distal finer-grained deposits offshore Montserrat (12–14). Fast and long-runout underwater volcaniclastic density currents Large volumes of pyroclastic material were delivered into the ocean during the eruption at Hunga volcano, creating dense underflows that were steered down its steep flanks. It is most likely that the pyroclastic material was predominantly derived from partial collapses of the eruption column, but we cannot fully preclude that currents may have been fed in part by other eruption processes, including jetting or fountaining. Initially, these volca- niclastic density currents would have been a dominantly gas-particle mixture, transitioning into a water-particle mixture as they cooled and mixed with seawater (23). We cannot dis- cern the precise extent of that transition, underlining our broader use of “volcaniclastic” rather than “pyroclastic” density current. Vol- caniclastic density currents were initially steered along preexisting relief into a valley 15 km southeast of the caldera. In this location, cur- rents intersected with the domestic cable side-on and were deflected to the north and south (i.e., parallel to the cable) by the topog- raphy. On the basis of the time between the first collapses of the eruptive column into the ocean (T0a and T0b) and severing of the domestic cable, we calculate a distance-averaged tran- sit (front) flow velocity of 63 to 122 km/hour (17.6 to 33.8 m/s; table S1). This observation is notable, given the inherent challenges in un- derwater monitoring, particularly during an ongoing large eruption. Despite the higher resistance provided by the surrounding seawater compared with air, the velocities of volcaniclastic density currents offshore Hunga volcano fall within the range measured for pyroclastic density currents on land (20) (table S4). The velocities at Hunga volcano are higher than previously documented for underwater density currents elsewhere in the ocean, including turbidity currents triggered by large-magnitude earthquakes, river floods, and oceanographic processes (Table 1 and Fig. 3). The fastest transit velocities for turbidity currents are up to 72 km/hour [20 m/s; based on cable breaks during the 1929 Grand Banks landslide (40) and the 2006 Pingtung earth- quake offshore Taiwan (37)]. The volcaniclastic density currents at Hunga volcano were steered into deeper water along tortuous paths created by irregular volcanic topography, where they severed the international cable 70 km from the volcano (Fig. 1). Assuming this current was the same one that also broke the domestic cable in- dicates a transit velocity of 32 to 51 km/hour (8.9 to 14.2 m/s). This is extraordinarily fast given the distance traveled. However, our data do not en- able us to determine how many currents were triggered at Hunga volcano. It is plausible that damage to the international cable resulted from currents triggered by eruption collapses con- siderably later in the eruption cycle (after T0a and conceivably after T0b), in which case these distal velocities are underestimates. What explains the fast current speeds? The sheer mass and manner of delivery of material to the ocean (i.e., direct and rapid vertical entry of a fast-collapsing plume) at Hunga volcano is distinct from other mecha- nisms of particulate density current generation, such as river plumes that enter the ocean later- ally, and landslide-triggered turbidity currents that initiate on far lower angle slopes and where material in the parent flow must first disintegrate and mix (Fig. 3). The dominantly downward rather than lateral trajectory toward and through the air-ocean boundary also makes the ocean-entry mechanism of dense pyroclas- tic material at Hunga volcano distinct from terrestrially initiated, ocean-entering pyro- clastic density currents such as those observed at Montserrat. The pyroclastic density currents that entered the ocean at Montserrat first trav- eled 4 km laterally over land, along the Tar River valley, after a dome collapse (12–14). In contrast, the formative mechanism at Hunga volcano is better described as a vertical jet or fountain collapse of a gas-particle mixture (53), wherein a huge sediment load of dense volcanic pyroclasts [up to 2.8 g/cm3 (52)] fell vertically from considerable height (several kilometers) as the eruption column collapsed into the ocean (Fig. 3). According to the modified Chézy equation [a simplified approach often used to model behavior of turbidity currents (52)], to main- tain the high current velocities observed at Hunga volcano would require a combination of a steep slope, thick current, and/or high sediment concentrations (expressed as the “depth averaged” value for a vertical profile through the current) (52, 54). Assuming pre- viously accepted basal and upper friction values for underwater density currents (54), to attain the high transit velocities on the edifice flanks requires current thicknesses on the order of tens of meters, with depth-averaged sediment concentrations of up to 5%, or else even-thicker currents (hundreds of meters thick) with lower depth-averaged sediment concentrations (1 to 2% concentration by volume). Near-bed sediment concentrations are likely substan- tially higher than these depth-averaged values (8, 55). These concentrations are particularly high for underwater particulate density cur- rents (55), as evidenced by the deposition of lobes on steep (up to ~10°) slopes of the edifice, implying a dense granular basal layer (41–43). Currents at Hunga volcano had sufficient inertia to flow upslope in some areas (and overtop relief of up to 860 m), such as south of the domestic cable break and to reach the international cable (figs. S4 and S5). This inertia was provided by their initial velocity and concentration and by additional entrained mass due to the substan- tial seafloor erosion. We conclude that fast ve- locities proximal to Hunga volcano result from (i) the potential energy generated from the di- rect, vertical entry of dense and large-volume pyroclastic fluxes into the ocean; (ii) the ex- tremely steep (up to 45°) submerged edifice flanks, which were the location of the impact zone for the collapsed material (53); and (iii) the additional mass gained by the currents as they eroded into the edifice flanks. Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 6 of 7 RES EARCH | R E S E A R C H A R T I C L E Implications for hazards assessment 13. J. Trofimovs, R. S. J. Sparks, P. J. Talling, Sedimentology 55, A push to enhance telecommunications links across the South Pacific and Caribbean will necessitate the crossing of volcanically active regions by new subsea cable routes with a need to assess threats posed to remote coastal com- munities and the international communica- tions infrastructure that serves them. This requires more-extensive seafloor mapping to identify submerged volcanoes that may ex- perience similar eruptions, while offshore monitoring, such as using fiber-optic sensing along telecommunications cables (56, 57), is required to provide early warning. We show that volcaniclastic density currents triggered when an eruption collapses into the ocean can maintain high densities over distances of >100 km and attain speeds up to 122 km/hour, providing a fundamentally new view of their behavior and associated hazards. We con- firm that bedforms observed on many other shallow submerged volcanoes worldwide can be produced by powerful eruptions, demon- strating that the hazards experienced at Hunga volcano can occur elsewhere (4, 8, 27). Explo- sive eruptions from these often unsurveyed and unmonitored submerged volcanoes can produce high-energy submarine density currents and warrant far greater consideration as tsunami- genic sources and as primary threats to vul- nerable coastal communities and critical subsea infrastructure. 617–634 (2008). 14. J. Trofimovs et al., Bull. Volcanol. 74, 391–405 (2012). 15. S. Seabrook et al., Research Square [Preprint] (2023); https://www.researchsquare.com/article/rs-2395332/v1. 16. C. Gomez, F. Lavigne, D. S. Hadmoko, N. Lespinasse, P. Wassmer, Geomorphology 110, 118–127 (2009). 17. D. E. Jessop et al., J. Volcanol. Geotherm. Res. 245–246, 81–97 (2012). 18. M. A. Hofton, E. Malavassi, J. B. Blair, Geophys. Res. Lett. 33, L21306 (2006). 19. P. J. Talling et al., Nat. Commun. 13, 4193 (2022). 20. L. Scharff, M. Hort, N. R. Varley, Sci. Rep. 9, 7386 (2019). 21. G. A. Valentine, J. Volcanol. Geotherm. Res. 87, 117–140 (1998). 22. I. Vilajosana, G. Khazaradze, E. Surinach, E. Lied, K. Kristensen, Cold Reg. Sci. Technol. 49, 2–10 (2007). 23. P. Kokelaar, P. Raine, M. J. Branney, Bull. Volcanol. 70, 23–54 (2007). 24. P. Kokelaar, S. Koniger, J. Geol. Soc. London 157, 517–536 (2000). 25. J. T. Gilchrist, A. M. Jellinek, E. E. Hooft, S. Wanket, Nat. Geosci. 16, 446–453 (2023). 26. E. L. Pope et al., Earth Planet. Sci. Lett. 493, 12–24 (2018). 27. D. Casalbore et al., Sedimentology 68, 1400–1438 (2021). 28. J. Karstens et al., Earth Planet. Sci. Lett. 616, 118215 (2023). 29. J. C. Borrero et al., Pure Appl. Geophys. 180, 1–22 (2023). 30. P. Lynett et al., Nature 609, 728–733 (2022). 31. K. Pakoksung, A. Suppasri, F. Imamura, Sci. Rep. 12, 15187 (2022). 32. P. Poli, N. M. Shapiro, Geophys. Res. Lett. 49, e2022GL098123 (2022). 33. S. R. Proud, A. T. Prata, S. Schmauß, Science 378, 554–557 (2022). 34. S. J. Purkis et al., Sci. Adv. 9, eadf5493 (2023). 35. D. A. Yuen et al., Earthq. Res. Adv. 2, 100134 (2022). 36. The World Bank and the Global Facility for Disaster Reduction and Recovery, “The January 15, 2022 Hunga Tonga-Hunga Ha’apai eruption and tsunami, Tonga: Global rapid post disaster damage estimation (grade) report” (International Bank for Reconstruction and Development/The World Bank, 2022); https://thedocs.worldbank.org/en/doc/ b69af83e486aa652d4232276ad698c7b-0070062022/ original/GRADE-Report-Tonga-Volcanic-Eruption.pdf. 37. L. Carter, R. Gavey, P. J. Talling, J. T. Liu, Oceanography 27, RE FE RENCES AND N OT ES 58–67 (2014). 1. G. Lube, E. C. Breard, T. Esposti-Ongaro, J. Dufek, B. Brand, Nat. Rev. Earth Environ. 1, 348–365 (2020). 2. E. C. Breard et al., Nat. Geosci. 9, 767–771 (2016). 3. E. C. Breard, G. Lube, Earth Planet. Sci. Lett. 458, 22–36 (2017). 4. C. E. J. de Ronde, G. J. Massoth, E. T. Baker, J. E. Lupton, in Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of Processes within the Earth, S. F. Simmons, I. Graham, Eds., Special Publication No. 10 (Society of Economic Geologists, 2003), pp. 91–109. 5. M. R. Auker, R. S. J. Sparks, L. Siebert, H. S. Crosweller, J. Ewert, J. Appl. Volcanol. 2, 2 (2013). 6. L. G. Mastin, J. B. Witter, J. Volcanol. Geotherm. Res. 97, 195–214 (2000). 7. B. Sovilla, M. Schaer, M. Kern, P. Bartelt, J. Geophys. Res. Earth Surf. 113, F01010 (2008). 38. J. Syvitski et al., Nat. Rev. Earth Environ. 3, 179–196 (2022). 39. M. Cassidy, L. Mani, Nature 608, 469–471 (2022). 40. B. C. Heezen, D. B. Ericson, M. Ewing, Deep Sea Res. 1, 193–202 (1954). 41. R. M. Iverson, Rev. Geophys. 35, 245–296 (1997). 42. R. M. Iverson, J. W. Vallance, Geology 29, 115–118 (2001). 43. J. J. Major, J. Geol. 105, 345–366 (1997). 44. C. J. Stevenson, C. A. L. Jackson, D. M. Hodgson, S. M. Hubbard, J. T. Eggenhuisen, J. Sediment. Res. 85, 1058–1081 (2015). 45. P. J. Talling et al., Nature 450, 541–544 (2007). 46. A. Manganet et al., J. Geophys. Res. 115, F03040 (2010). 47. O. Korup et al., Earth Planet. Sci. Lett. 261, 578–589 (2007). 48. M. J. Cartigny, D. Ventra, G. Postma, J. H. van Den Berg, Sedimentology 61, 712–748 (2014). 8. E. L. Pope et al., Sci. Adv. 8, eabj3220 (2022). 9. B. Sovilla, J. N. McElwaine, M. Y. Louge, C. R. Phys. 16, 97–104 49. W. O. Symons, E. J. Sumner, P. J. Talling, M. J. Cartigny, M. A. Clare, Mar. Geol. 371, 130–148 (2016). (2015). 10. R. S. J. Sparks, H. Sigurdsson, S. N. Carey, J. Volcanol. Geotherm. Res. 7, 87–96 (1980). 11. S. Carey, H. Sigurdsson, C. Mandeville, S. Bronto, in Volcanic Hazards and Disasters in Human Antiquity, Floyd W. McCoy, C. Heiken, Eds., Special Paper 345 (Geological Society of America, 2000), pp. 1–14. 12. J. Trofimovs et al., Geology 34, 549–552 (2006). 50. O. E. Sequeiros, J. Geophys. Res. Oceans 117, C04003 (2012). 51. J. J. Mountjoy et al., Sci. Adv. 4, eaar3748 (2018). 52. See supplementary materials. 53. G. A. Valentine, Bull. Volcanol. 82, 20 (2020). 54. K. Konsoer, J. Zinger, G. Parker, J. Geophys. Res. Earth Surf. 118, 216–228 (2013). 57. T. Nishimura et al., Sci. Rep. 11, 6319 (2021). 58. D. J. Piper, B. Savoye, Sedimentology 40, 557–582 (1993). 59. P. Colantoni et al., Giornale di Geologia 54, 190–213 (1992). 60. B. C. Heezen, M. Ewing, AAPG Bull. 39, 2505–2514 (1955). 61. M. S. Heijnen et al., Nat. Commun. 13, 3448 (2022). 62. D. McClung, P. Shaerer, The Avalanche Handbook (Mountaineers Books, ed. 3, 2006). 63. K. Mackay, Digital Elevation Models of Hunga Volcano; pre- and post- 15 January 2022 eruption, version 1.0, Zenodo (2022); https://doi.org/10.5281/zenodo.7456324. AC KNOWLED GME NTS We thank the Kingdom of Tonga for allowing us to undertake this research and the Nippon Foundation-GEBCO Seabed 2030 project and their alumni for their support. This work would not have been possible without the captain, crew, and scientists aboard RV Tangaroa (Voyage TAN2206) and the use of SEA-KIT International’s USV Maxlimer for mapping the caldera. We extend our gratitude to everyone involved in these voyages and for assistance in processing the results. We thank B. Sugar of South Sea Charters, Nuku’alofa, for providing the photographs of the eruption plume that appear in fig. S6. The staff of the British Ocean Sediment Core Research Facility (BOSCORF), including C. McGuire, R. Garnett, M. Charidemou, and M. Edwards, are thanked for supporting MSCL-CIS data collection and for providing logistical support to receive and store cores in the UK. We thank R. Williams and two anonymous reviewers for their constructive comments and suggestions. Funding: This work was supported by Natural Environment Research Council grant NE/X00239X/1 (M.A.C. and I.A.Y.); Natural Environment Research Council grant NE/X003272/1 (J.E.H., M.A.C., and I.A.Y.); Natural Environment Research Council grant NE/X002454/1 (D.T.); International Cable Protection Committee (M.A.C. and I.A.Y.); and the Nippon Foundation grant: NIWA-Nippon Foundation Tonga Eruption Seabed Mapping Project (S.W., R.W., S.S., K.M., E.L., and M.W.). Author contributions: Conceptualization: M.A.C., I.A.Y., P.J.T., E.P., K.M., R.W., S.W., T.K., S.H., C.d.R., M.U., S.K., S.F., S.P., D.V., R.R., and V.K. Methodology: K.M., M.W., I.A.Y., M.A.C., J.E.H., S.F., S.P., S.W., S.S., P.J.T., and E.P. Investigation: S.S., S.W., K.M., M.W., M.A.C., I.A.Y., J.E.H., S.C., and M.R. Visualization: M.A.C., I.A.Y., S.W., J.E.H., and S.C. Funding acquisition: M.A.C., I.A.Y., J.E.H., M.W., K.M., and D.T. Project administration: M.A.C., M.W., and K.M. Writing – original draft: M.A.C. and I.A.Y. Writing – review & editing: S.W., R.W., S.S., K.M., J.E.H., E.L., P.J.T., E.P., S.C., M.R., T.K., D.T., S.H., C.d.R., M.U., S.K., S.F., S.P., D.V., R.R., V.K., and M.W. Competing interests: M.A.C. is the marine environmental adviser to the International Cable Protection Committee. The other authors do not have any competing interests. Data and materials availability: Core logs, photographs, and coordinates are provided in the methods section of the supplementary materials. The pre- and post-eruption bathymetric data can be accessed in Zenodo (63). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/ about/science-licenses-journal-article-reuse. This research was funded in whole or in part by the Natural Environment Research Council (NE/X00239X/1, NE/X003272/1, NE/X002454/1), a cOAlition S organization. The author will make the Author Accepted Manuscript (AAM) version available under a CC BY public copyright license. SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adi3038 Materials and Methods Figs. S1 to S6 Tables S1 to S5 References (64–89) 55. C. K. Paull et al., Nat. Commun. 9, 4114 (2018). 56. P. Jousset et al., Nat. Commun. 13, 1753 (2022). Submitted 17 April 2023; accepted 1 August 2023 10.1126/science.adi3038 Clare et al., Science 381, 1085–1092 (2023) 8 September 2023 7 of 7
10.1126_science.adj5438
RES EARCH PHOTONICS Ultrafast mode-locked laser in nanophotonic lithium niobate Qiushi Guo1,2,3*, Benjamin K. Gutierrez4, Ryoto Sekine1, Robert M. Gray1, James A. Williams1, Luis Ledezma1,5, Luis Costa1, Arkadev Roy1, Selina Zhou1, Mingchen Liu1, Alireza Marandi1,4* Mode-locked lasers (MLLs) generate ultrashort pulses with peak powers substantially exceeding their average powers. However, integrated MLLs that drive ultrafast nanophotonic circuits have remained elusive because of their typically low peak powers, lack of controllability, and challenges when integrating with nanophotonic platforms. In this work, we demonstrate an electrically pumped actively MLL in nanophotonic lithium niobate based on its hybrid integration with a III-V semiconductor optical amplifier. Our MLL generates ~4.8-ps optical pulses around 1065 nm at a repetition rate of ~10 GHz, with energies exceeding 2.6 pJ and peak powers beyond 0.5 W. The repetition rate and the carrier-envelope offset frequency of the output can be controlled in a wide range by using the driving frequency and the pump current, providing a route for fully stabilized on-chip frequency combs. M ode-locked lasers (MLLs), which gen- erate intense and coherent ultrashort optical pulses on picosecond and femto- second timescales, have enabled nu- merous sciences and technologies in photonics such as extreme nonlinear optics (1), supercontinuum generation (2), optical atomic clocks (3), optical frequency combs (4), biological imaging (5), and photonic computing (6). Today’s state-of-the-art MLLs are based on discrete fiber- based and free-space optical components and are expensive, power demanding, and bulky. Realizing MLLs on integrated photonic plat- forms promises widespread use of ultrafast photonic systems that are currently limited to table-top laboratory experiments. However, the performance of integrated MLLs has not been on par with their table-top counterparts, lacking the required peak intensities and de- grees of controllability required for on-chip ultrafast optical systems (7). A major challenge lies in the simultaneous realization of large laser gain and an efficient mode-locking mech- anism on integrated photonic platforms. Al- though III-V semiconductor gain media can be electrically pumped and exhibit a very high gain per unit length and high saturation powers (8), the conventional method of achieving mode- locking and short pulse generation on the same semiconductor chip requires a narrow range of pumping current, thus substantially limiting the output power and the tunability of the in- tegrated MLLs (9, 10). To realize high-peak-power integrated MLLs, a promising approach consists of the hybrid integration of a semiconductor gain medium and an external mode-locking element based on electrooptic (EO) or nonlinear optical effects. Thin-film lithium niobate (TFLN) has emerged as a promising integrated nonlinear photonic platform with access to power-efficient and high-speed EO modulation (11, 12) and strong quadratic [c(2)] optical nonlinearity (13, 14). 1Department of Electrical Engineering, California Institute of Technology, Pasadena, CA, USA. 2Photonics Initiative, Advanced Science Research Center, City University of New York, NY, USA. 3Physics Program, Graduate Center, City University of New York, New York, NY, USA. 4Department of Applied Physics, California Institute of Technology, Pasadena, CA, USA. 5Jet Propulsion Laboratory, Pasadena, CA, USA. *Corresponding author. Email: qguo@gc.cuny.edu (Q.G.); marandi@caltech.edu (A.M.) A D Reflector RF signal Output coupler Turning point B Intracavity phase gain PM Laser cavity length (L) 0 C Amplitude gain f m f0 … f -n Turning point loss … f n t f RF signal 50 Ω load III-V gain chip Reflector ground signal ground Loop mirror TFLN SiO2 Silicon Fig. 1. Principle and design of integrated actively MLL laser. (A) Diagram of active mode-locking through intracavity phase modulation. (B) Illustration of mode- locking in the time domain. (C) Illustration of mode-locking in the frequency domain. (D) Schematic of the integrated actively MLL. Guo et al., Science 382, 708–713 (2023) 10 November 2023 1 of 6 RES EARCH | R E S E A R C H A R T I C L E Hybrid integration of semiconductor gain with TFLN enables a strong interplay between the laser gain and the EO or nonlinear effects to achieve active or passive mode-locking with high efficiency and tunability. Moreover, many of the nonlinear and ultrafast optical function- alities such as supercontinuum generation (15), optical parametric oscillation (16–18), pulse shortening (19), all-optical switching (20), and quantum squeezing (21) can be realized in quasi–phase-matched LN nanophotonic devices with orders-of-magnitude lower peak powers compared with other platforms. Therefore, de- veloping high-peak-power MLLs integrated into nanophotonic LN can enable a suite of nonlinear and ultrafast optical phenomena on a chip, promising integrated photonic sys- tems with unprecedented performance and functionalities. In this work, we demonstrate a high-peak-power, electrically pumped, inte- grated, actively MLL by hybrid integration of III-V semiconductors and LN nanophotonics. Our MLL exploits the high laser gain of III-V semiconductors and the efficient active optical phase modulation in LN nanophotonic wave- guides as the mode-locking mechanism. Such a design eliminates the complexities associated with realizing gain and saturable absorption on the same semiconductor chip, allowing a much higher output power and a wider tunability of the laser. Mode-locked laser operating principle and design Figure 1A shows the concept of active mode- locking by intracavity EO phase modulation. In the time domain, when a phase modulator (PM) is driven by a sinusoidal radio frequency (RF) signal at a frequency fm, the intracavity phase modulation is equivalent to the cavity length modulation. Therefore, the laser cavity can be considered as having a moving end mirror with a sinusoidal motion at frequency fm. When an optical signal inside the cavity strikes this moving end mirror and gets reflected back, its optical frequency acquires a Doppler shift. After successive round trips, these Doppler shifts will accumulate, resulting in no steady- state solution. However, when a short circulat- ing pulse strikes the end mirror at either of the “turning points” where the mirror reverses its direction (the extremum of the phase varia- tion as shown in Fig. 1B), it will not acquire a Doppler frequency shift, but instead, a small quadratic phase modulation or chirp (22). Thus, a steady-state optical pulse can be maintained in the laser cavity after successive round trips. Although in principle, optical pulses can occur at either of the two phase modulation extrema and acquire chirps of different signs, the dis- persion in the cavity can compensate for the chirp imposed by the PM at one extremum, and further chirp the pulse formed at another extremum (23). The mode-locking mechanism can also be understood in the frequency do- main (Fig. 1C). When the intracavity phase modulation frequency fm matches the cavity free spectral range (FSR), the sidebands pro- duced by each of the running axial modes are injected into the adjacent axial modes, result- ing in the phase locking of adjacent modes. Notably, in MLLs, these modes will lase be- cause of the presence of laser gain within the cavity, whereas in EO comb sources, they are generated by dispersing the energy from a single pump laser line (24–27). Figure 1D shows the design of our integrated MLL based on this principle. In our MLL, an electrically pumped single-angled facet (SAF) GaAs gain chip is butt-coupled to a TFLN chip that contains an integrated EO PM and a broadband loop mirror. A Fabry-Perot laser cavity configuration is formed between the reflective facet on the left end of the SAF gain chip and the broadband loop mirror on the TFLN chip. Here, an integrated PM is preferred over a Mach Zehnder interferometer (MZI)– C D Ground Signal Ground 10 µm 50 µm 5 µm A |E| 1.0 B E ) m µ ( Y 2 0 -2 F 2 µm SiO2 SiO2 0 |E| 1.0 H -8 -4 0 X (µm) 4 8 -8 -4 0 X (µm) 4 8 0.0 G III-V gain chip TFLN chip 50 µm 500 µm Fig. 2. Integrated actively MLL laser on TFLN. (A) Cross-sectional view of the PM region and the distribution of microwave field (white arrows) at 10 GHz and the optical field of the fundamental TE mode (color map) at 1065 nm. The RF electrodes are marked in yellow. |E|, normalized electric field. (B) False-colored SEM image of the PM region in the fabricated device. The RF electrodes are marked in yellow and the optical waveguide is marked in blue. (C) SEM image of the broadband loop mirror. (D) The curved coupling region of the broadband loop mirror. (E and F) The fundamental TE mode profile at 1065 nm in both the (E) SAF gain chip waveguide and (F) TFLN waveguide taper. (G) Optical microscope image showing the coupling region between the two chips. (H) Dark-field optical microscope image of the integrated MLL when operating. Guo et al., Science 382, 708–713 (2023) 10 November 2023 2 of 6 RES EARCH | R E S E A R C H A R T I C L E A B C D Fig. 3. Characterization of integrated actively MLL. (A) Schematic of the experimental setup. CTL, continuously tunable CW laser; FPD, fast photodetector; OSA, optical spectrum analyzer; ESA, electrical spectrum analyzer; YDFA, ytterbium-doped fiber amplifier. (B) The optical spectrum of the MLL output as a function of the RF driving frequency fm. (C) Intensity autocorrelation of the MLL output as a function of the fm. (D) Laser average output power versus Idrive when fm = 10.17 GHz. Ith, threshold driving current. based intensity modulator (IM) because the PM offers a lower insertion loss and avoids effects from the dc bias drift of the MZI mod- ulator (28). For the PM, we used an RF coplanar waveguide (CPW) design with ground-signal- ground (GSG) configuration to ensure the transmission of the RF wave with minimal radiative loss. Characterization of mode-locked lasers We fabricated our devices on a 700-nm-thick X-cut magnesium oxide (MgO)–doped TFLN on a SiO2/Si substrate. In the PM region (Fig. 2A), the RF CPW was fabricated on top of the SiO2 cladding layer. Such a design allowed us to achieve high modulation efficiency (simulated VpL = 1.1 V(cid:1)cm) by having a small gap (4 mm here) between the ground and signal electrodes and a significant overlap between the RF field and the optical field in the waveguide (12, 29). We designed the geometry of the RF CPW to ensure a 50-W impedance around 10 GHz. Figure 2, B to D, shows the scanning electron microscope (SEM) images of the fabricated PM and the loop mirror. We adopted a curved coupling-region design in the loop mirror to ensure broadband reflection (see supplemen- tary materials section I for details). Based on the length (1.5 mm) and the refractive index of the SAF gain chip around 1065 nm, we es- timate that a ∼3-mm-long TFLN waveguide in- cluding the loop mirror section can lead to a laser cavity FSR of ∼10 GHz. Figure 2E shows the 1065-nm fundamental transverse electric (TE)–mode profile in the wave- guide of the SAF gain chip. To minimize the coupling loss between the SAF gain chip and the TFLN chip, the top width of the input facet of the TFLN waveguide was tapered out to be 10.3 mm. The 1065-nm fundamental TE-mode profile in the tapered TFLN waveguide is shown in Fig. 2F. This design ensures a maximal over- lap with the optical mode produced by the SAF gain chip. We estimate a chip coupling loss of 3.4 to 3.9 dB in our experiments (see supple- mentary materials section II for details). The chip coupling loss can be further reduced by using a mode-size converter based on an in- verted taper edge coupler embedded in a poly- mer waveguide (30, 31). Figure 2G shows the microscope image of the chip coupling region after the alignment. When the SAF gain chip is electrically pumped with a driving current (Idrive) of 160 mA, we observed green light (the second harmonic of the 1065-nm light) inside the laser cavity (Fig. 1H), which indicated a high intracavity power around 1065 nm and a good alignment between the two chips. We characterized the integrated actively MLL using the optical setup shown in Fig. 3A. We applied a ~280-mW sinusoidal RF signal to the left end of the CPW of the PM using the RF probe. The right end of the CPW is terminated Guo et al., Science 382, 708–713 (2023) 10 November 2023 3 of 6 RES EARCH | R E S E A R C H A R T I C L E Reference f 1 f 2 … A r e w o P B … Optical frequency D ) m B d ( r e w o P E f m=10.17 GHz RF on RF off f 3dB=0.35 nm 0 -10 -20 -30 -40 -50 -60 1060 1062 1064 1066 Wavelength (nm) 1068 1070 f m=10.17 GHz 4.81 ps 8 6 4 2 W/ Pulse shaper W/O Pulse shaper Gaussian fit Gaussian fit 5.03 ps . ) . u a ( n o i t l a e r r o c o t u A 0 -60 -50 -40 -30 -20 -10 0 Time (ps) 10 20 30 40 50 60 C ) m B d ( r e w o P -40 -60 -80 -100 -120 3 f 1=3.68 GHz y0 = -1.92456E-10, xc = 3.68371E9 w = 4115238.93752, A = 0.02817 H = 4.35836E-9 4) W n ( 2 r e w o P 3.95 MHz 0 3.67 3.68 3.69 3.70 RF frequency (GHz) 25 20 15 10 5 0 ) W n ( r e w o P f2=6.49 GHz y0 = -9.2114E-10, xc = 6.49094E9 w = 4074667.13986, A = 0.11225 H = 1.75378E-8 f m=10.17 GHz 3.91 MHz 6.48 6.50 6.49 RF frequency (GHz) 4 5 6 7 8 9 10 RF frequency (GHz) Fig. 4. Finding the mode-locking regime of integrated MLL. (A) Illustration of the generation of heterodyne beat notes. (B) Evolution of heterodyne beat notes with the fm. The mode-locking regime is marked by the white dashed box. (C) Heterodyne beat notes measured at fm = 10.17 GHz. (Insets) Zoomed-in view of the two beat notes at f1 = 3.68 GHz and f2 = 6.49 GHz. Blue symbols represent measured data and solid red curves indicate Lorentz fits. (D) Output optical spectra of the MLL when the RF drive at 10.17 GHz is on (red) and off F 7) s p ( i h t d w e s u P l 6 5 4 P-0.125 RF reference 100 200 PRF (mW) 300 400 500 (blue). The side lobes around 1062.7 and 1067 nm can be due to the reflections in the gap between the gain chip and the TFLN chip. The fast modulation can be due to the reflections within the gain chip. (E) Intensity autocorrelation traces of the MLL output measured at fm = 10.17 GHz with (red) and without (blue) the external pulse shaper. Symbols represent measurement data and solid curves indicate Gaussian fits. a.u., arbitrary units. (F) Dependence of pulse width on PRF. The red dashed line represents the 1= scaling law according to the HME. p 8 ffiffiffiffiffiffiffi PRF by another RF probe with a commercially avail- able 50-W load resistor. While the gain chip was pumped with an Idrive of 185.2 mA, we simul- taneously collected the laser output spectra, the intensity autocorrelation of the laser output, and the heterodyne beat notes between two neighboring laser emission lines and a narrow- linewidth (~10 kHz) reference continuous wave (CW) tunable laser. As shown in Fig. 3B, when we scanned the fm, the laser output exhibited a clear spectral broadening between 10.1 and 10.4 GHz (labeled by the white dashed box). Within this fm range, two distinct intensity autocorrelation peaks separated by ~98 ps emerged (Fig. 3C), indicating that optical pulses are formed. At an fm of 10.17 GHz, we measured the laser output power from the output facet of the TFLN chip. As shown in Fig. 3D, the laser exhibits a low threshold Idrive of 22 mA. Given the measured coupling loss of ~11 dB between the TFLN waveguide and the single-mode lensed fiber, the on-chip laser out- put average power is more than 50 mW when the Idrive is greater than 180 mA. Heterodyne beat notes were used to char- acterize the mode-locking and resulting fre- quency comb. As illustrated in Fig. 4A, when the frequency of the reference CTL is resting in be- tween the two neighboring comb lines of the MLL near the center of its spectrum, two RF beat notes at f1 and f2 are generated on the fast detector. As shown in Fig. 4B, when fm is be- tween 10.165 and 10.173 GHz, as labeled by the white dashed box, two spectrally narrow beat notes are observed. This suggests that within this range of fm, the laser operates in the mode- locked regime, producing ultrashort optical pulses with high coherence. When fm is slightly detuned between 10.165 and 10.173 GHz, f1 and f2 can shift significantly with fm (Fig. 4B). This indicates that the carrier frequency of the MLL sensitively depends on fm. However, when the fm is further detuned from the cavity FSR, the MLL exhibits a transition to a turbulent re- gime (32), which is manifested by multiple noisy beat notes around f1 and f2 in Fig. 4B. In the turbulent regime, the laser can still emit ultrashort pulses as shown in Fig. 3C, albeit with low coherence. As shown in Fig. 4C, at fm = 10.17 GHz, we obtained two spectrally narrow RF beat notes at f1 = 3.68 GHz and f2 =6.49 GHz, with full width at half maximum (FWHM) linewidths of 3.95 and 3.91 MHz, respectively. Given that the RF drive has a very small phase noise, and no active locking of the laser cavity was used here, the linewidths of the heterodyne beat notes can be mainly limited by the drift of pulse carrier frequency. As shown in Fig. 4D, when a 280-mW RF drive at 10.17 GHz was applied to the PM, significant spectral broad- ening was observed. The pulse spectrum was centered at 1064.9 nm and the FWHM of the spectrum was 0.35 nm. Meanwhile, the inten- sity autocorrelation trace (Fig. 4E) indicated that the MLL produced one strong pulse at one of the modulation turning points, where- as the other pulse was significantly suppressed. We fit the intensity autocorrelation trace with a Gaussian function becasue active mode lock- ing produces Gaussian pulses according to the Haus master equation (HME) (33). The fitting yielded a pulse width of 4.81 ps with the ex- ternal pulse shaping, and 5.03 ps without. Be- cause the pulse shaper can compensate for the chirp on the output pulse and the additional chirp imposed by the fibers and the YDFA, we expect the output pulse width directly after the Guo et al., Science 382, 708–713 (2023) 10 November 2023 4 of 6 RES EARCH | R E S E A R C H A R T I C L E A 220 200 180 160 140 120 100 80 60 ) A m ( e v i r d I C ) m B d ( r e w o P -60 -80 -100 -120 Power (dBm) B 0 -10 -20 -30 -40 -50 ) A m ( e v i r d I 220 200 180 160 140 120 100 80 60 Autocorrelation (a.u.) 9 8 7 6 5 4 3 -20 -10 0 10 20 Time (ps) 194.6 mA 193.6 mA 192.6 mA 191.6 mA 190.6 mA 189.6 mA D ) z H G 10.20 ( m f 10.15 m u m i t p O 10.10 10.05 1055 1060 1065 1070 Wavelength (nm) 2 4 RF frequency (GHz) 6 8 10 140 160 I 180 drive (mA) 200 Fig. 5. Current tuning of integrated actively MLL. (A) The optical spectrum of the MLL output as a function of Idrive. (B) Autocorrelation trace of the MLL output as a function of Idrive. a.u., arbitrary units. (C) Tuning of the heterodyne beat notes by the Idrive. In (A) to (C), fm is fixed at 10.18 GHz. (D) Dependence of optimum fm for mode-locking on Idrive. MLL facet to be between 4.81 ps and 5.03 ps. The pulse width of 4.81 ps after pulse shaping cor- responds to a time-bandwidth product of 0.445, which is very close to the transform-limited time-bandwidth product (0.44) of a Gaussian pulse (34). To conservatively estimate the pulse energy and peak power, we used the measured output average power of 53 mW at Idrive = 185.2 mA and assumed that both pulses exist in the cavity. Hence, the output pulse energy of our MLL was at least 2.6 pJ and the pulse peak power was greater than 0.51 W. 8 p We further studied the pulse width limits of our MLL. First, we found that the measured pulse only slightly decreased when the RF driving power PRF was increased (Fig. 4F), ffiffiffiffiffiffiffiffi which is in good agreement with the 1= PRF scaling law (red dashed line) according to the HME (35). We also found that further increas- ing the RF power will not shorten the pulse significantly. Instead, it can lead to laser in- stability due to RF heating. The HME with cavity group velocity dispersion, neglecting non- linear effects within the laser cavity, predicts a minimum pulse width of ~2.5 ps (36). The ex- perimentally measured pulse width was wider, likely due to several factors that are not cap- tured by the HME such as the complex carrier dynamics and two-stage gain recovery in the III-V gain medium (37) and gain bandwidth narrowing from intracavity etalon (22) or hole burning effects (38) (see supplementary mate- rials section IV for detailed analysis). Electrical tuning of mode-locked lasers The Idrive can serve as an important tuning knob of our MLL. Figure 5, A and B, shows the de- pendence of the output spectra and autocor- relation of the MLL on the Idrive with 280-mW RF drive fixed at 10.17 GHz. Within a wide range of Idrive (140 to 205 mA), optical pulses can be formed inside the laser. In addition, the carrier frequency of the MLL blueshifted by ~0.3 nm as the Idrive was increased from 140 to 200 mA (Fig. 5A). This was likely caused by the blueshift of the peak wavelength of the gain spectrum owing to band filling and screening effects induced by carrier injection (39). We also investigated the effect of Idrive on the coherence property and the frep of the laser. We kept the RF drive fixed at 280 mW and 10.18 GHz. As the Idrive was tuned from 189.6 to 194.6 mA, the laser transitioned from the turbulent to the mode-locked regime, and then back to the turbulent regime (Fig. 5C). These results suggest that, with a frequency- stable reference CW laser and active feedback on Idrive, it may be possible to lock the carrier frequency of the MLL and operate the device as a stable frequency comb, as the repetition rate (frep) of the MLL has already been locked by the external RF oscillator. As shown in Fig. 5D, when we widely varied Idrive from 144 to 204 mA, the optimum fm that enables mode- locking with high coherence could be varied from 10.04 to 10.23 GHz, indicating that the frep of the laser could also be adjusted by ~200 MHz. Moreover, the optimum fm increased almost linearly with Idrive, which resulted from an increase of the cavity FSR caused by carrier injection in the gain medium. Although fur- ther increasing the Idrive would not lead to laser output power saturation, we did not observe pulse formation at higher Idrive beyond 205 mA. This is likely attributed to the significant de- tuning of the cavity FSR or the instability of laser mode-locking due to more pronounced self-phase modulation in the gain medium (37). Conclusions and outlook We have demonstrated an integrated actively MLL in nanophotonic lithium niobate oper- ating around 1065 nm, which offers the high- est output pulse energy and peak power of any integrated MLLs in nanophotonic platforms (fig. S9 and table S2). Our MLL allows for a wide tuning range of the laser frep of ~200 MHz and precise control of the laser’s coherence properties. The current tuning capability of our MLL indicates that active feedback to Idrive can achieve simultaneous locking of the carrier frequency and frep of the MLL. This allows the MLL to operate as a stable fre- quency comb with locked carrier frequency offset (fCEO) and frep. Finally, seamless integra- tion of our high–peak power MLL with other c(2) nonlinear optical functionalities provided by TFLN offers opportunities for realizing new integrated photonic systems, such as fully integrated supercontinuum sources, self-referenced frequency combs, and atomic clocks. REFERENCES AND NOTES 1. M. Wegener, Extreme Nonlinear Optics: An Introduction 2. (Springer Science & Business Media, 2005). J. M. Dudley, G. Genty, S. Coen, Rev. Mod. Phys. 78, 1135–1184 (2006). 3. A. D. Ludlow, M. M. Boyd, J. Ye, E. Peik, P. O. Schmidt, Rev. Mod. Phys. 87, 637–701 (2015). 4. T. Udem, R. Holzwarth, T. W. Hänsch, Nature 416, 233–237 (2002). 5. D. W. Piston, Trends Cell Biol. 9, 66–69 (1999). 6. A. Marandi, Z. Wang, K. Takata, R. L. Byer, Y. Yamamoto, Nat. Photonics 8, 937–942 (2014). 7. M. L. Davenport, S. Liu, J. E. Bowers, Photon. Res. 6, 468 (2018). 8. P. J. Delfyett et al., IEEE J. Quantum Electron. 28, 2203–2219 (1992). 9. A. Hermans et al., APL Photonics 6, 096102 (2021). 10. E. Vissers et al., Opt. Express 30, 42394–42405 (2022). 11. C. Wang et al., Nature 562, 101–104 (2018). 12. M. Xu et al., Optica 9, 61 (2022). 13. C. Wang et al., Optica 5, 1438 (2018). 14. J. Lu et al., Optica 6, 1455 (2019). 15. M. Jankowski et al., Optica 7, 40 (2020). Guo et al., Science 382, 708–713 (2023) 10 November 2023 5 of 6 RES EARCH | R E S E A R C H A R T I C L E 16. L. Ledezma et al., Sci. Adv. 9, eadf9711 (2023). 17. J. Lu et al., Optica 8, 539 (2021). 18. T. P. McKenna et al., Nat. Commun. 13, 4532 (2022). 19. A. Roy et al., Nat. Photonics 16, 162–168 (2022). 20. Q. Guo et al., Nat. Photonics 16, 625–631 (2022). 21. R. Nehra et al., Science 377, 1333–1337 (2022). 22. D. Kuizenga, A. Siegman, IEEE J. Quantum Electron. 6, 694–708 (1970). 23. R. Nagar, D. Abraham, G. Eisenstein, Opt. Lett. 17, 1119–1121 (1992). 24. M. Zhang et al., Nature 568, 373–377 (2019). 25. M. Yu et al., Nature 612, 252–258 (2022). 26. Y. Hu et al., Nat. Photonics 16, 679–685 (2022). 27. A. Rueda, F. Sedlmeir, M. Kumari, G. Leuchs, H. G. L. Schwefel, Nature 568, 378–381 (2019). 28. J. P. Salvestrini, L. Guilbert, M. Fontana, M. Abarkan, S. Gille, J. Lightwave Technol. 29, 1522–1534 (2011). 29. M. Jin, J. Chen, Y. Sua, P. Kumar, Y. Huang, Opt. Lett. 46, 1884–1887 (2021). 30. C. Hu et al., Opt. Express 29, 5397–5406 (2021). 31. E. Vissers, S. Poelman, C. O. de Beeck, K. Van Gasse, B. Kuyken, Opt. Express 29, 15013–15022 (2021). 32. F. Kärtner, D. Zumbühl, N. Matuschek, Phys. Rev. Lett. 82, 4428–4431 (1999). 33. U. Keller, R. Paschotta, Ultrafast Lasers (Springer, 2021). 34. A. E. Siegman, Lasers (University Science Books, 1986). 35. H. A. Haus, IEEE J. Sel. Top. Quantum Electron. 6, 1173–1185 (2000). 36. Q. Guo, data and computer code for “Ultrafast mode-locked laser in nanophotonic lithium niobate”, version 2, Figshare (2023); https://doi.org/10.6084/m9.figshare.24208434. 37. G. P. Agrawal, N. A. Olsson, IEEE J. Quantum Electron. 25, 2297–2306 (1989). 38. A. Uskov, J. Mork, J. Mark, IEEE J. Quantum Electron. 30, 1769–1781 (1994). 39. S. Schmitt-Rink, D. S. Chemla, D. A. Miller, Phys. Rev. B Condens. Matter 32, 6601–6609 (1985). ACKN OWLED GMEN TS The device nanofabrication was performed at the Kavli Nanoscience Institute (KNI) at Caltech. The authors thank K. Vahala for loaning equipment. Q.G. thanks M.Xu for the helpful discussions. Funding: The authors acknowledge support from ARO grant no. W911NF-23-1-0048, NSF grant nos. 1846273 and 1918549, AFOSR award FA9550-23-1-0755, and NASA JPL. The authors thank NTT Research for their financial and technical support. Author contributions: Q.G. and A.M. conceived the project. Q.G. fabricated the devices with assistance from R.S. Q.G. performed the measurements and numerical simulation. R.S., J.W., B.K.G., R.M.G., L.L., L.C., and S.Z. assisted with the measurements. B.K.G., R.M.G., A.R., and M.L. helped with the numerical simulation and data analysis. Q G. wrote the manuscript with input from all authors. A.M. supervised the project. Competing interests: Q.G. and A.M. are inventors on a patent application (US patent application no. 17/500,425) that covers the concept and implementation of the actively MLL in this work. L.L. and A.M. are involved in developing photonic integrated nonlinear circuits at PINC Technologies Inc. L.L. and A.M. have an equity interest in PINC Technologies Inc. The remaining authors declare no competing interests. Data and materials availability: All data are available in the manuscript or the supplementary materials. The data files supporting the plots in the main text and the computer code for simulating the MLL are available at Figshare (36). License information: Copyright © 2023 the authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original US government works. https://www.science.org/about/science- licenses-journal-article-reuse SUPPLEMENTARY MATERIALS science.org/doi/10.1126/science.adj5438 Materials and Methods Supplementary Text Figs. S1 to S9 Tables S1 and S2 References (40–48) Submitted 2 July 2023; accepted 4 October 2023 10.1126/science.adj5438 Guo et al., Science 382, 708–713 (2023) 10 November 2023 6 of 6