The full dataset viewer is not available (click to read why). Only showing a preview of the rows.
The dataset generation failed
Error code:   DatasetGenerationError
Exception:    ArrowInvalid
Message:      JSON parse error: Missing a closing quotation mark in string. in row 3
Traceback:    Traceback (most recent call last):
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/packaged_modules/json/json.py", line 145, in _generate_tables
                  dataset = json.load(f)
                File "/usr/local/lib/python3.9/json/__init__.py", line 293, in load
                  return loads(fp.read(),
                File "/usr/local/lib/python3.9/json/__init__.py", line 346, in loads
                  return _default_decoder.decode(s)
                File "/usr/local/lib/python3.9/json/decoder.py", line 340, in decode
                  raise JSONDecodeError("Extra data", s, end)
              json.decoder.JSONDecodeError: Extra data: line 2 column 1 (char 49851)
              
              During handling of the above exception, another exception occurred:
              
              Traceback (most recent call last):
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/builder.py", line 1995, in _prepare_split_single
                  for _, table in generator:
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/packaged_modules/json/json.py", line 148, in _generate_tables
                  raise e
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/packaged_modules/json/json.py", line 122, in _generate_tables
                  pa_table = paj.read_json(
                File "pyarrow/_json.pyx", line 308, in pyarrow._json.read_json
                File "pyarrow/error.pxi", line 154, in pyarrow.lib.pyarrow_internal_check_status
                File "pyarrow/error.pxi", line 91, in pyarrow.lib.check_status
              pyarrow.lib.ArrowInvalid: JSON parse error: Missing a closing quotation mark in string. in row 3
              
              The above exception was the direct cause of the following exception:
              
              Traceback (most recent call last):
                File "/src/services/worker/src/worker/job_runners/config/parquet_and_info.py", line 1529, in compute_config_parquet_and_info_response
                  parquet_operations = convert_to_parquet(builder)
                File "/src/services/worker/src/worker/job_runners/config/parquet_and_info.py", line 1154, in convert_to_parquet
                  builder.download_and_prepare(
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/builder.py", line 1027, in download_and_prepare
                  self._download_and_prepare(
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/builder.py", line 1122, in _download_and_prepare
                  self._prepare_split(split_generator, **prepare_split_kwargs)
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/builder.py", line 1882, in _prepare_split
                  for job_id, done, content in self._prepare_split_single(
                File "/src/services/worker/.venv/lib/python3.9/site-packages/datasets/builder.py", line 2038, in _prepare_split_single
                  raise DatasetGenerationError("An error occurred while generating the dataset") from e
              datasets.exceptions.DatasetGenerationError: An error occurred while generating the dataset

Need help to make the dataset viewer work? Make sure to review how to configure the dataset viewer, and open a discussion for direct support.

text
string
meta
dict
\section{Introduction} \label{intro} Star clusters are a powerful tool in the investigation of Galaxy structure and dynamics, star formation and evolution processes, and as observational constraints to N-body codes. This applies especially to the long-lived and populous globular clusters (GCs) that, because of their relatively compact nature, can be observed in most regions of the Galaxy, from near the center to the remote halo outskirts. In general terms, the structure of most star clusters can be described by a rather dense core and a sparse halo, but with a broad range in the concentration level. In this context, the standard picture of a GC assumes a isothermal central region and a tidally truncated outer region (e.g. \citealt{Binney1998}). Old GCs, in particular, can be virtually considered as dynamically relaxed systems (e.g. \citealt{NoGe06}). During their lives clusters are continually affected by internal processes such as mass loss by stellar evolution, mass segregation and low-mass star evaporation, and external ones such as tidal stress and dynamical friction e.g. from the Galactic bulge, disk and giant molecular clouds (e.g. \citealt{Khalisi07}; \citealt{Lamers05}; \citealt{GnOs97}). Over a Hubble time, these processes tend to decrease cluster mass, which may accelerate the core collapse phase for some clusters (\citealt{DjMey94}, and references therein). Consequently, these processes, combined with the presence of a central black hole (in some cases) and physical conditions associated to the initial collapse, can affect the spatial distribution of light (or mass) both in the central region and at large radii (e.g. \citealt{GLO99}; \citealt{NoGe06}). It is clear from the above that crucial information related to the early stages of Galaxy formation, and to the cluster dynamical evolution, may be imprinted in the present-day internal structure and large-scale spatial distribution of GCs (e.g. \citealt{MvdB05}; \citealt{GCProp}). To some extent, this reasoning can be extended to the open clusters (OCs), especially the young, which are important to determine the spiral arm and disk structures and the rotation curve of the Galaxy (e.g. \citealt{Friel95}; \citealt{DiskProp}). Consequently, the derivation of reliable structural parameters of star clusters, GCs in particular, is fundamental to better define their parameter space. This, in turn, may result in a deeper understanding of the formation and evolution processes of the star clusters themselves and the Galaxy. Three different approaches have been used to derive structural parameters of star clusters. The more traditional one is based on the surface-brightness profile (SBP), which considers the spatial distribution of the brightness of the component stars, usually measured in circular rings around the cluster center. The compilation of Harris (1996, and the 2003 update\footnote{\em http://physun.physics.mcmaster.ca/Globular.html}) presents a basically uniform set of parameters for 150 Galactic GCs. Among their structural parameters, the core (\mbox{$\rm R_c$}), half-light (\mbox{$\rm R_{hL}$}) and tidal (\mbox{$\rm R_t$}) radii, as well as the concentration parameter $c=\log(\mbox{$\rm R_t$}/\mbox{$\rm R_c$})$, were based mostly on the SBP database of \citet{TKD95}. SBPs do not necessarily require cluster distances to be known, since the physically relevant information contained in them is essentially related to the relative brightness of the member stars. In principle, it is easy to measure integrated light. However, SBPs are more efficient near the cluster center than in the outer parts, where noise and background starlight may be a major contributor. Another potential source of noise is the random presence of bright stars, either from the field or cluster members, especially outside the central region in the less-populous GCs or most of the OCs. Structural parameters derived from such SBPs would certainly be affected. One way to minimise this effect is the use of wide rings throughout the whole radius range, but this would cause spatial resolution degradation on the profiles, especially near the center. The obvious alternative to SBPs is to use star counts to build radial density profiles (RDPs), in which only the projected number-density of stars is taken into account, regardless of the individual star brightness. This technique is particularly appropriate for the outer parts, provided a statistically significant, and reasonably uniform, comparison field is available to tackle the background contamination. On the other hand, contrary to SBPs, RDPs are less efficient in central regions of populous clusters where the density of stars (crowding) may become exceedingly large. In such cases it may not be possible to resolve individual stars with the available technology. Finally, a more physically significant profile can be built by mapping the cluster's stellar mass distribution, which essentially determines the gravitational potential and drives most of the dynamical evolution. However, mass density profiles (MDPs) not only are affected by the same technical problems as the RDPs but, in addition, the cluster distance, age and a reliable mass-luminosity relation are necessary to build them. In principle, the three kinds of profiles are expected to yield different values for the structural parameters under similar photometric conditions, since each profile is sensitive to different cluster parameters, especially the age and dynamical state. Qualitatively, the following effects, basically related to dynamical state, can be expected. Large-scale mass segregation drives preferentially low-mass stars towards large radii (while evaporation pushes part of these stars beyond the tidal radius, into the field), and high-mass stars towards the central parts of clusters. If the stellar mass distribution of an evolved cluster can be described by a spatially variable mass function (MF) flatter at the cluster center than in the halo, the resulting RDP (and MDP) radii should be larger than SBP ones. The differences should be more significant for the core than the half and tidal radii, since the core would contain, on average, stars more massive than the halo and especially near the tidal radius. Besides, the presence of bright stars preferentially in the central parts of young clusters (\citealt{DetAnalOCs} and references therein) should as well lead to smaller SBP core and half-light radii than the respective RDP ones. Another relevant issue is related to depth-limited photometry. When applied to the observation of objects at different distances, depth-limited photometry samples stars with different brightness (or mass), especially at the faint (or low-mass) end. Thus, it would be interesting to quantify the changes produced in the derived parameters when RDPs, MDPS and SBPs are built with depth-limited photometry, as well as to check how the structural parameters derived from one type of profile relate to the equivalent radii measured in the other profiles. In the present work we face the above issues by deriving structural parameters of star clusters built under controlled conditions, in which the radial distribution of stars follows a pre-established analytical profile, and field stars are absent. Effects introduced by mass segregation (simulated by a spatially variable mass function), age and structure are also considered. This work focuses on profiles built in the near-infrared range. The main goal of the present work is to examine relations among structural parameters measured in the different radial profiles, built under ideal conditions, especially noise-free photometry and as small as possible statistical uncertainties (using a large number of stars). In this sense, the results should be taken as upper-limits. \begin{table*} \caption[]{Model star cluster specifications} \label{tab1} \renewcommand{\tabcolsep}{2.65mm} \renewcommand{\arraystretch}{1.2} \begin{tabular}{ccccccccrcccc} \hline\hline Model&$R_t/R_c$&c&$\chi_0$&$\chi_t$&Age&\mbox{$\rm [Fe/H]$}&$m_i$&$m_s$&$\langle m\rangle$ &$\rm M_J(TO)$&$\rm M_J(bright)$&$\rm M_J(faint)$\\ & & & & &(Myr)&&(\mbox{$\rm M_\odot$})&(\mbox{$\rm M_\odot$})&(\mbox{$\rm M_\odot$})&(mag)&(mag)&(mag)\\ (1) &(2) &(3) &(4) &(5) &(6) &(7) &(8) &(9) &(10)&(11) &(12) &(13)\\ \hline GC-A &5&0.7&0.00&1.35&$10^4$&$-1.5$&0.15&1.02&0.43&$+2.86$&$-2.14$&$+9.12$\\ GC-B&20&1.3&0.00&1.35&$10^4$&$-1.5$&0.15&1.02&0.43&$+2.86$&$-2.14$&$+9.12$\\ GC-C&20&1.3&0.00&0.00&$10^4$&$-1.5$&0.15&1.02&0.46&$+2.86$&$-2.14$&$+9.12$\\ GC-D&40&1.6&0.00&1.35&$10^4$&$-1.5$&0.15&1.02&0.43&$+2.86$&$-2.14$&$+9.12$\\ OC-A&15&1.2&0.30&1.35&$10^3$&$~~0.0$&0.15&2.31&0.59&$+0.32$&$-2.68$&$+9.18$\\ OC-B&15&1.2&0.30&1.35&$100$&$~~0.0$&0.15&5.42&0.92&$-1.82$&$-4.82$&$+9.18$\\ OC-C&15&1.2&0.30&1.35&$10$&$~~0.0$&0.15&18.72&1.76&$-4.82$&$-8.82$&$+9.18$\\ \hline \end{tabular} \begin{list}{Table Notes.} \item Col.~3: concentration parameter $c=\log(\mbox{$\rm R_t$}/\mbox{$\rm R_c$})$. Cols.~4 and 5: mass function slopes at the cluster center and tidal radius. Cols.~8-10: lower, upper and average star mass. Col.~11: absolute J magnitude at the turnoff (TO). Cols.~12 and 13: absolute J magnitude at the bright and faint ends. \end{list} \end{table*} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig1.eps}} \caption{Model star cluster specifications. Panel (a): a random selection of $n$ in the range $0\leq n\leq1$ produces King-like RDPs in the range $0\leq R\leq\mbox{$\rm R_t$}$ (see Eq.~\ref{eq2}). Panel (b): Radially-variable mass function slopes $\left(\frac{dN}{dm}\propto m^{-(1+\chi)}\right)$ used in the models. Panel (c): Padova isochrones used to simulate the mass-luminosity relation of the star cluster models. The 10\,Gyr, $\mbox{$\rm [Fe/H]$}=-1.5$ metallicity isochrone is adopted in the globular cluster models. Panel (d): distribution of concentration parameters of the GCs in H03 with peaks at $c\approx1.6,~1.3,~{\rm and}~0.7$. Panel (e): model fraction of stars brighter than $M_J=M_{J_{TO}}+\Delta_{TO}$. In all cases, the fraction of stars brighter than the TO ($M_{J_{TO}}$) is below the $1\%$ level.} \label{fig1} \end{figure} This work is structured as follows. In Sect.~\ref{ModelSCs} we present the star cluster models and build radial profiles with depth-limited photometry. In Sect.~\ref{Struc} we derive structural parameters from each profile, discuss their dependence on depth, and compare similar radii derived from the different types of profiles. In Sect.~\ref{N6397} we compare relations derived from model parameters with those of the nearby GC NGC\,6397. Concluding remarks are given in Sect.~\ref{Conclu}. \section{The model star clusters} \label{ModelSCs} For practical reasons, the model star clusters are simulated by first establishing the number-density radial distribution. The approach we follow is to build star clusters of different ages and concentration parameters, with the spatial distribution of stars truncated at the tidal radius (\mbox{$\rm R_t$}). Stars are distributed with distances to the cluster center in the range $0\leq R\leq\mbox{$\rm R_t$}$, with the $R$ coordinate having a number-frequency given by a function similar to a \citet{King62} three-parameter surface-brightness profile. The mass and brightness of each star are subsequently computed according to a pre-defined mass function and mass-luminosity relation consistent with the model age. The last step is required for the derivation of the MDP and SBPs. We point out that different, more sophisticated analytical models have also been used to fit the SBPs of Galactic and extra-Galactic GCs, other than \citet{King62} profile. The most commonly used are the single-mass, modified isothermal sphere of \citet{King66} that is the basis of the Galactic GC parameters given by \citet{TKD95} and H03, the modified isothermal sphere of \citet{Wilson75}, that assumes a pre-defined stellar distribution function (which results in more extended envelopes than \citealt{King66}), and the power-law with a core of \citet{EFF87} that has been fit to massive young clusters especially in the Magellanic Clouds (e.g. Mackey \& Gilmore 2003a,b,c). Each function is characterised by different parameters that are somehow related to the cluster structure. However, the purpose here is not to establish a ``best'' fitting function of the structure of star clusters in general. Instead, we want to quantify changes in the structural parameters, derived from RDPs, MDPs and SBPs of star clusters with the stellar distribution assumed to follow an analytical function, under different photometric conditions. We expect that changes in a given parameter should have a small dependence, if any at all, on the adopted functional form. The adopted King-like radial distribution function is expressed as \begin{equation} \label{eq1} \frac{dN}{2\pi\,R\,dR}=\sigma_0\left[\frac{1}{\sqrt{1+(R/R_c)^2}} - \frac{1}{\sqrt{1+(R_t/R_c)^2}}\right]^2, \end{equation} where $\sigma_0$ is the projected number-density of stars at the cluster center, and \mbox{$\rm R_c$}\ and \mbox{$\rm R_t$}\ are the core and tidal radii, respectively. Since structural differences are basically controlled by the ratio $\mbox{$\rm R_t$}/\mbox{$\rm R_c$}$, we set $\mbox{$\rm R_c$}=1$ in all models. Such a King-like RDP (for $\sigma_0=1.0$) is obtained by numerically inverting the relation (see App.~\ref{Transf}) \begin{equation} \label{eq2} n(R) = \frac{x^2-4u(\sqrt{1+x^2}-1)+u^2\ln(1+x^2)}{u^2\ln{u^2}-(u-1)(3u-1)}, \end{equation} where $x\equiv R/R_c$ and $u^2\equiv 1+(R_t/R_c)^2$. Thus, a random selection of numbers in the range $0\leq n\leq1$ produces a King-like radial distribution of stars with the radial coordinate in the range $0\leq R/R_t\leq1$. The $R/R_t$ curves as a function of $n$ for the models considered in this work are shown in Fig.~\ref{fig1} (Panel a). Once a given star has been assigned a radial coordinate, its mass is computed with a probability proportional to the mass function \begin{equation} \label{eq3} \frac{dN}{dm}\propto m^{-(1+\chi)}, \end{equation} where the slope varies with $R$ according to $\chi=\chi(R)=\chi_t + (\chi_t-\chi_0)(R/R_t-1)$, where $\chi_0$ and $\chi_t$ are the mass function slopes at the cluster center and tidal radius, respectively (Table~\ref{tab1} and Fig.~\ref{fig1}). Thus, the presence of large-scale mass segregation in a star cluster can be characterised by a slope $\chi_0$ flatter than $\chi_t$. Mass values distributed according to Eq.~\ref{eq3} are obtained by randomly selecting numbers in the range $0\leq n\leq1$ and using them in the relation of mass with $n~\rm{and~} \chi$ (App.~\ref{Transf}) \begin{equation} \label{eq4} m=\left\{ \begin{array}{lc} m_i\,(m_s/m_i)^n, & \rm{for~\chi=0.0,}\\ m_s/[(1-n)(m_s/m_i)^\chi+n]^{1/\chi}, & \rm{otherwise}, \end{array} \right . \end{equation} where $m_i$ and $m_s$ are the lower and upper mass values considered in the models (Table~\ref{tab1}). In what follows we adopt the 2MASS\footnote{\em http://www.ipac.caltech.edu/2mass/releases/allsky/} photometric system to build SBPs. Finally, the 2MASS \mbox{$\rm J$}, \mbox{$\rm H$}\ and \mbox{$\rm K_s$}\ magnitudes for each star are obtained according to the mass-luminosity relation taken from the corresponding model (Table~\ref{tab1}) Padova isochrone (\citealt{Girardi02}). For illustrative purposes the model isochrones are displayed in Fig.~\ref{fig1} (panel c). The set of models considered here is intended to be objectively representative of the star cluster parameter space. For globular clusters we use the standard age of 10\,Gyr and the spatially uniform metallicity $\mbox{$\rm [Fe/H]$}=-1.5$, which is typical of the metal-poor Galactic GCs (e.g. \citealt{GCProp}). However, we note that abundance variations have been suggested to occur within GCs (e.g. \citealt{Gratton04}). Basically, small to moderate metallicity gradients would produce slight changes in the colour and magnitude of the stars in different parts of the cluster, which has no effect on the (star-count derived) RDPs and MDPs. The effect on the SBPs may be small as well, provided that the magnitude bin used to build the SBPs is wide enough to accommodate such magnitude changes. As for the core/tidal structure we consider the ratios $R_t/R_c=40,~20,~15,~{\rm and}~5$, or equivalently the concentration parameters $c=\log{(R_t/R_c)\approx1.6,~1.3,~1.2,~{\rm and}~0.7}$, which roughly correspond to the peaks in the distribution of $c$ values presented by the regular (non-post core collapse) GCs given in H03 (Fig.~\ref{fig1}, panel d). Models GC-A, B and D take into account mass segregation by means of a flat ($\chi_0=0.00$) mass function at the center and a Salpeter (1955) IMF ($\chi_t=1.35$) at the tidal radius. GC-C model is similar to GC-B, except that it considers a uniform, heavily depleted MF ($\chi_0=0.00$) throughout the cluster. OCs are represented by solar-metallicity models with the ages 10\,Myr (to allow for the presence of bright stars in young OCs), 100\,Myr (somewhat evolved OCs) and 1\,Gyr (intermediate-age OCs), $R_t/R_c=15$ ($c\approx1.2$) and a spatially variable MF (Table~\ref{tab1}). The values of $c$ and the core/halo MF slopes are representative of OCs (\citealt{DetAnalOCs}). Another effect not considered here is differential absorption. In principle, low to moderate differential absorption should have a minimum effect on the radial profiles, because of the same reasons as those given above for the metallicity gradient. High values, on the other hand, would affect RDPs as well, because of a radially-dependent loss of stars due to depth-limited photometry. However, inclusion of this effect is beyond the scope of the present work. As expected, the fraction of stars brighter than the turnoff (TO) in the resulting star cluster models is significantly smaller than 1\% (Fig.~\ref{fig1}, panel e). Thus, we had to use a total number of stars of $1\times10^9$ in all models, so that the radial profiles resulted statistically significant (small $1\sigma$ Poisson error bars) especially at the shallowest magnitude depth. \subsection{Depth-varying radial profiles} \label{DeptVP} The radial profiles were built considering all stars brighter than a given magnitude threshold, with the TO as reference. At the bright end, statistically significant GC profiles were obtained for $\mbox{$\rm \Delta_{TO}$}\equiv M_{J,th}-M_{J,TO}=-5$, where $M_{J,th}$ and $M_{J,TO}$ are the threshold and TO absolute magnitudes in the 2MASS \mbox{$\rm J$}\ band. At the faint end, GC-models have $\mbox{$\rm \Delta_{TO}$}=6.3$. OC models have $\mbox{$\rm \Delta_{TO}$}=-3~{\rm and~} -4$ at the bright end, and $\mbox{$\rm \Delta_{TO}$}=8.9,~11.0,~{\rm and}~14.0$, at the faint end. Starting at the bright magnitude end, RDPs, MDPs and SBPs were built considering stars with the \mbox{$\rm J$}\ magnitude brighter than a given faint threshold, with the magnitude depth increasing in steps of $\mbox{$\rm \Delta_{TO}$}=1$, up to the respective faint magnitude end. Figure~\ref{fig2} displays a selection of profiles corresponding to both extremes in magnitude depths, for the GC-D and OC-C models. These profiles are representative of the whole set of models, especially in terms of the small uncertainties associated with each radial coordinate. Reflecting the large differences in the number of stars at different photometric depths, the central values of the number and mass densities, and surface-brightness, vary significantly from the shallowest to the deepest profiles. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig2.eps}} \caption{A selection of RDPs (top panels), MDPs (middle) and 2MASS \mbox{$\rm J$}\ magnitude SBPs (bottom) that illustrate structural changes under different magnitude depths. Arbitrary units (au) are used both for the radial coordinate and projected area.} \label{fig2} \end{figure} \begin{table*} \caption[]{Model star cluster structural parameters for different photometric depths} \label{tab2} \renewcommand{\tabcolsep}{1.3mm} \renewcommand{\arraystretch}{1.2} \begin{tabular}{cccccccccccc} \hline\hline &\multicolumn{3}{c}{RDP}&&\multicolumn{3}{c}{MDP}&&\multicolumn{3}{c}{SBP (\mbox{$\rm J$}\ band)}\\ \cline{2-4}\cline{6-8}\cline{10-12} $\Delta_{TO}$&\mbox{$\rm R_c$}&\mbox{$\rm R_{hSC}$}&\mbox{$\rm R_t$}&&\mbox{$\rm R_c$}&\mbox{$\rm R_{hM}$}&\mbox{$\rm R_t$}&&\mbox{$\rm R_c$}&\mbox{$\rm R_{hL}$}&\mbox{$\rm R_t$}\\ (mag)&(au)&(au)&(au)&&(au)&(au)&(au)&&(au)&(au)&(au)\\ (1) &(2) &(3) &(4) &&(5) &(6) &(7) &&(8) &(9) &(10) \\ \hline &\multicolumn{11}{c}{Model: GC-A; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=5.0$}\\ \cline{2-12} $-5.0$&$0.78\pm(\dag)$&$1.02\pm(\dag)$&$4.61\pm0.01$&&$0.78\pm(\dag)$&$1.02\pm(\dag)$&$4.61\pm0.01$&&$0.75\pm0.01$&$1.01\pm(\dag)$&$4.80\pm0.01$\\ $~0.0$&$0.76\pm(\dag)$&$1.02\pm(\dag)$&$4.77\pm0.01$&&$0.76\pm(\dag)$&$1.02\pm(\dag)$&$4.77\pm0.01$&&$0.75\pm0.01$&$1.01\pm(\dag)$&$4.79\pm0.01$\\ $+6.3$&$1.00\pm(\dag)$&$1.19\pm(\dag)$&$5.00\pm0.01$&&$0.92\pm(\dag)$&$1.14\pm(\dag)$&$4.91\pm(\dag)$&&$0.75\pm0.01$&$1.03\pm(\dag)$&$4.80\pm0.01$\\ \hline &\multicolumn{11}{c}{Model: GC-B; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=20.0$}\\ \cline{2-12} $-5.0$&$0.87\pm0.01$&$2.03\pm(\dag)$&$17.31\pm0.08$&&$0.87\pm0.01$&$2.03\pm(\dag)$&$17.31\pm0.08$&&$0.86\pm0.01$&$2.04\pm0.01$&$17.82\pm0.05$\\ $~0.0$&$0.83\pm0.01$&$2.03\pm(\dag)$&$18.72\pm0.08$&&$0.83\pm0.01$&$2.03\pm(\dag)$&$18.72\pm0.08$&&$0.86\pm0.01$&$2.03\pm(\dag)$&$17.80\pm0.03$\\ $+6.3$&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$20.00\pm(\dag)$&&$0.95\pm(\dag)$&$2.27\pm(\dag)$&$19.28\pm0.03$&&$0.86\pm0.01$&$2.05\pm(\dag)$&$17.80\pm0.02$\\ \hline &\multicolumn{11}{c}{Model: GC-C; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=20.0$}\\ \cline{2-12} $-5.0$&$1.00\pm(\dag)$&$2.38\pm0.01$&$20.02\pm0.03$&&$1.00\pm(\dag)$&$2.38\pm0.01$&$20.02\pm0.04$&&$1.00\pm(\dag)$&$2.38\pm0.01$&$19.94\pm0.06$\\ $~0.0$&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$20.00\pm0.01$&&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$20.00\pm0.01$&&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$19.95\pm0.03$\\ $+6.3$&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$20.00\pm(\dag)$&&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$20.00\pm(\dag)$&&$1.00\pm(\dag)$&$2.39\pm(\dag)$&$19.97\pm0.03$\\ \hline &\multicolumn{11}{c}{Model: GC-D; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=40.0$}\\ \cline{2-12} $-5.0$&$0.90\pm0.01$&$2.81\pm0.02$&$33.96\pm0.19$&&$0.90\pm0.01$&$2.81\pm0.02$&$33.96\pm0.19$&&$0.91\pm0.01$&$2.82\pm(\dag)$&$34.18\pm0.05$\\ $~0.0$&$0.86\pm0.01$&$2.82\pm(\dag)$&$37.15\pm0.17$&&$0.86\pm0.01$&$2.82\pm(\dag)$&$37.15\pm0.17$&&$0.91\pm0.01$&$2.82\pm(\dag)$&$34.00\pm0.05$\\ $+6.3$&$1.00\pm(\dag)$&$3.30\pm(\dag)$&$39.99\pm0.01$&&$0.96\pm(\dag)$&$3.14\pm(\dag)$&$38.51\pm0.07$&&$0.91\pm0.01$&$2.82\pm(\dag)$&$34.20\pm0.04$\\ \hline &\multicolumn{11}{c}{Model: OC-A; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=15.0$}\\ \cline{2-12} $-3.0$&$0.82\pm0.01$&$1.70\pm(\dag)$&$12.85\pm0.07$&&$0.82\pm0.01$&$1.70\pm(\dag)$&$12.85\pm0.07$&&$0.81\pm0.01$&$1.72\pm(\dag)$&$13.18\pm0.02$\\ $~0.0$&$0.78\pm0.01$&$1.72\pm(\dag)$&$13.78\pm0.06$&&$0.78\pm0.01$&$1.72\pm(\dag)$&$13.78\pm0.06$&&$0.82\pm0.01$&$1.72\pm(\dag)$&$13.19\pm0.02$\\ $+8.9$&$1.00\pm(\dag)$&$2.08\pm(\dag)$&$15.00\pm0.01$&&$0.91\pm(\dag)$&$1.93\pm(\dag)$&$14.43\pm0.03$&&$0.81\pm0.01$&$1.73\pm(\dag)$&$13.20\pm0.01$\\ \hline &\multicolumn{11}{c}{Model: OC-B; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=15.0$}\\ \cline{2-12} $-3.0$&$0.72\pm0.01$&$1.61\pm(\dag)$&$13.30\pm0.08$&&$0.72\pm0.01$&$1.61\pm(\dag)$&$13.30\pm0.08$&&$0.76\pm0.01$&$1.61\pm(\dag)$&$12.75\pm0.03$\\ $~0.0$&$0.70\pm0.02$&$1.61\pm(\dag)$&$13.67\pm0.08$&&$0.70\pm0.02$&$1.61\pm(\dag)$&$13.67\pm0.08$&&$0.77\pm0.01$&$1.61\pm(\dag)$&$12.74\pm0.03$\\ $+11.0$&$1.00\pm(\dag)$&$2.08\pm(\dag)$&$15.00\pm(\dag)$&&$0.84\pm0.01$&$1.84\pm(\dag)$&$14.33\pm0.04$&&$0.74\pm0.02$&$1.63\pm(\dag)$&$12.75\pm0.02$\\ \hline &\multicolumn{11}{c}{Model: OC-C; Input RDP parameters: $\mbox{$\rm R_c$}=1.0$, $\mbox{$\rm R_t$}=15.0$}\\ \cline{2-12} $-4.0$&$0.62\pm0.02$&$1.49\pm(\dag)$&$13.06\pm0.10$&&$0.62\pm0.02$&$1.49\pm(\dag)$&$13.05\pm0.10$&&$0.71\pm0.01$&$1.49\pm(\dag)$&$11.99\pm0.03$\\ $~0.0$&$0.62\pm0.02$&$1.49\pm(\dag)$&$13.10\pm0.10$&&$0.62\pm0.02$&$1.49\pm(\dag)$&$13.09\pm0.10$&&$0.71\pm0.01$&$1.49\pm(\dag)$&$11.99\pm0.03$\\ $+14.0$&$1.00\pm(\dag)$&$2.08\pm(\dag)$&$15.00\pm(\dag)$&&$0.70\pm0.02$&$1.70\pm(\dag)$&$14.35\pm0.05$&&$0.64\pm0.03$&$1.49\pm(\dag)$&$12.00\pm0.02$\\ \hline \end{tabular} \begin{list}{Table Notes.} \item ($\dag$): uncertainty smaller than 0.01 arbitrary units (au). The half-type radii are half-star counts (\mbox{$\rm R_{hSC}$}), half-mass (\mbox{$\rm R_{hM}$}) and half-light (\mbox{$\rm R_{hL}$}). \end{list} \end{table*} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig3.eps}} \caption{Structural parameters of the GC models. Top panels: Ratio of the tidal radius measured in profiles with a photometric depth \mbox{$\rm \Delta_{TO}$}\ with respect to that derived from the deepest one, for the RDPs (left panels), MDPs (vertical-middle) and SBPs (right). Horizontal-middle panels: half-type radii. Bottom: core radii. TO values are indicated by the dotted line. Except for GC-C (uniform mass function), the remaining models present changes in radii in the RDPs and MDPs. SBP radii are essentially uniform.} \label{fig3} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig4.eps}} \caption{Same as Fig.~\ref{fig3} for the OC models. For comparison purposes, the y-scale is the same as in Fig.~\ref{fig3}. Similarly to the GC models, radii changes are conspicuous in the RDPs and MDPs. } \label{fig4} \end{figure} \section{Structural parameters {\em vs.} photometry depth} \label{Struc} The depth-varying model SBPs are fit with the empirical three-parameter function introduced by \cite{King62} to describe the surface-brightness distribution of GCs, which is characterised by the presence of the core and tidal radii. For RDPs and MDPs we use the King-like analytical profile that describes the projected number-density of stars as a function of \mbox{$\rm R_c$}\ and \mbox{$\rm R_t$}, $\sigma(R)=\frac{dN}{2\pi\,R\,dR}$, as given by eq.~\ref{eq1}. We also compute the distances from the center which contains half of the cluster's total light, stars and mass. The half-star count (\mbox{$\rm R_{hSC}$}), light (\mbox{$\rm R_{hL}$}) and mass (\mbox{$\rm R_{hM}$}) radii are derived by directly integrating the corresponding profiles. A selection of the resulting structural parameters as a function of \mbox{$\rm \Delta_{TO}$}\ is given in Table~\ref{tab2}. For simplicity we only present the values obtained from the bright and faint magnitude ranges, as well as for $M_J\leq M_{J,TO}$. The whole set of parameters are contained in Figs.~\ref{fig3} - \ref{fig6}. At first glance, RDP and MDP radii present a significant decrease for shallower photometry, with respect to the intrinsic values. SBP radii, on the other hand, are more uniform. The most noticeable feature is that, except for GC-C (uniform mass function), RDP and MDP radii tend to become increasingly larger than SBP ones with increasing photometric depth. \subsection{Dependence on photometric depth} \label{DependDepth} In Fig.~\ref{fig3} we compare the radii measured in GC profiles built with a given photometric depth (e.g. $\mbox{$\rm R_c$}(\Delta_{TO})$) with the intrinsic ones, i.e. those derived from the deepest profiles ($R_{c,deep}$). RDP parameters are more affected than the MDP ones, while the SBP ones are essentially uniform, thus insensitive to photometric depth. Among the radii, RDP and MDP core are the most affected (underestimated), followed by the half and tidal radii. In the most concentrated model (GC-A, $c\approx0.7$), measurements or \mbox{$\rm R_c$}\ in the RDP may be underestimated by a factor $\approx25\%$ in profiles shallower than near the TO, with respect to $R_{c,deep}$, and $\approx20\%$ in MDPs. The effect is smaller in \mbox{$\rm R_{hSC}$}\ and \mbox{$\rm R_{hM}$}, which may be underestimated by $\approx15\%$ in the same profiles. The underestimation in the tidal radii is smaller than $\approx10\%$. As expected, RDP, MDP and SBP radii do not change when the mass function is uniform (GC-C model). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig5.eps}} \caption{GC model profiles. Ratio between the same type of radii as measured in RDPs and MDPs (left panels) and RDPs and SBPs (right panels). From top to bottom: tidal, half and core radii. TO values are indicated by the dotted line.} \label{fig5} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig6.eps}} \caption{Same as Fig.~\ref{fig5} for the OC models. For comparison purposes, the y-scale is the same as in Fig.~\ref{fig5}.} \label{fig6} \end{figure} Similar radii ratios in the OC models are examined in Fig.~\ref{fig4}. Qualitatively, the same conclusions drawn from the GC models apply to the OC ones. However, the underestimation factor of RDP radii increases for younger ages, to the point that \mbox{$\rm R_c$}\ drops to $\approx60\%$ of the deepest value for all profiles shallower than $\approx3$\, mag below the TO in the OC-C model ($10$\,Myr), and to $\approx70\%$ for OC-B ($100$\,Myr). The respective half-star count radii are affected by similar, although smaller, underestimation factors. MDP radii are less affected by cluster age than RDP ones. Similarly to the GC models (Fig.~\ref{fig3}), the three types of SBP radii are essentially insensitive to photometric depth, within uncertainties. We note that the presence of bright stars in the central region of young clusters (OC-C) appears to introduce a small dependence of the core radius on photometric depth (bottom-right panel). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig7.eps}} \caption{Top panels: relation of the half-type radii with the concentration parameter, for the RDPs (left panel), MDPs (middle) and SBPs (right). For each model, $R_h$ values increase for deeper profiles. Dashed line in panels (a) and (b): $R_h\sim c^2$. In panel (c): $\mbox{$\rm R_{hL}$}\sim c$. Bottom panels: concentration parameter as a function of photometric depth.} \label{fig7} \end{figure} \subsection{Comparison of similar radii among different profiles} \label{CompDifProf} Differences on the same type of radii among the profiles, introduced essentially by a spatially variable MF, are discussed in Fig.~\ref{fig5} for the GC models. Regardless of the model assumptions, RDP and MDP radii are essentially the same, except for the profiles corresponding to deep photometry, for which the RDP radii become slightly larger than the MDP ones. This occurs basically because of the larger fraction of low-mass stars at the outer parts of the clusters. Since all stars have equal weight in the building of the RDPs, the accumulation of low-mass stars at large radii ends up broadening the RDPs with respect to the MDPs. On the other hand, RDP core and half-star count radii tend to be larger than the SBP ones for profiles including stars fainter than near the TO. RDP \mbox{$\rm R_t$}\ may be 10 -- 20\% larger than SBP ones for all depths. As discussed above, the uniformly-depleted MF of GC-C model produces profiles whose radii are independent of photometry depth. The RDP to SBP core and tidal radii ratios decrease with concentration parameter. The RDP to SBP half-type radii ratios do not depend on $c$. The same analysis applied to the OC models is discussed in Fig.~\ref{fig6}. The presence of massive stars in young clusters enhances the RDP to MDP radii ratios, especially the core and to some extent, the half-type radii. This occurs for profiles that contain stars brighter than $\approx4$\,mag below the TO. For the youngest model (OC-C), the core radius measured in the RDP may be $\approx40\%$ larger than the MDP one. This effect is enhanced when RDP radii are compared to SBP ones, again decreasing in intensity from the core to tidal radii. For OC-C, RDP core, half and tidal radii are $\approx55\%$, $\approx40\%$, and $\approx25\%$ larger than the equivalent SBP ones. Comparing with the GC models (Fig.~\ref{fig5}), the presence of a larger fraction of more massive (brighter) stars towards the center in young clusters tend to enhance radii ratios of RDP with respect to MDP, and especially, RDP to SBP. \subsection{Further relations} \label{FurtRel} The models discussed in previous sections can be used as well to examine the dependence of the half-type radii with the concentration parameter, and to test how $c$ varies with photometric depth. These issues are presented in Fig.~\ref{fig7}. As already suggested by Figs.~\ref{fig3} and \ref{fig4}, the relation of the half radius with $c$, in a given model, changes significantly with photometric depth in RDPs (panel a) and MDPs (panel b). In SBPs, on the other hand, it is almost insensitive to depth (panel c). From eq.~\ref{eq1}, the half-star count radius is tightly related to the concentration parameter according to $\mbox{$\rm R_{hSC}$}=(0.69\pm0.01)+(1.01\pm0.01)\,c^2$. This curve fits well the values measured in the deepest RDP of all GC and OC models alike (panel a). Such a relation fails for the shallower profiles. A similar, but poorer, relation applies to the values derived from the deepest MDPs (panel b), $\mbox{$\rm R_{hM}$}=(0.63\pm0.09)+(0.99\pm0.05)\,c^2$. It fails especially for the young (OC) models. The GC SBPs, on the other hand, can be poorly fit with the linear function $\mbox{$\rm R_{hL}$}=(-0.9\pm0.1)+(2.4\pm0.1)\,c$ (panel c). Concentration parameters measured in RDPs and MDPs (panels d and e) change with photometric depth. Around the TO they reach the maximum value, which corresponds to a star cluster $\approx15\%$ more concentrated than the pre-established value (Table~\ref{tab1}). At the shallowest profiles $c$ presents a value intermediate between the maximum and the pre-established one, which is retrieved at the deepest profiles with the inclusion of the numerous low-mass stars. The exception again is the uniform MF model GC-C, whose $c$ values do not change with $\Delta_{TO}$. $c$ values measured in SBPs are essentially insensitive to photometric depth (panel f). \section{NGC\,6397: a test case} \label{N6397} We compare the results derived for the model star clusters with similar parameters measured in the $\mbox{$\rm M_V$}=-6.63$, nearby GC ($\mbox{$\rm d_\odot$}=2.3$\,kpc) NGC\,6397. Being populous is important to produce statistically significant radial profiles, while the proximity allows a few magnitudes fainter than the giant branch to be reached with depth-limited photometry. NGC\,6397 is a post-core collapse GC with evidence of large-scale mass segregation, as indicated by a mass function flatter at the center than outwards (\citealt{Andreuzzi04} and references therein). Additional relevant data (from H03) for the metal-poor ($\mbox{$\rm [Fe/H]$}=-1.95$) GC NGC\,6397 are the Galactocentric distance $\mbox{$\rm R_{GC}$}=6$\,kpc, half-light and tidal radii (measured in the V band) $\mbox{$\rm R_{hL}$}=2.33\arcmin$ and $\mbox{$\rm R_t$}=15.81\arcmin$, and Galactic coordinates $\ell=338.17^\circ$, $b=-11.96^\circ$. Thus, bulge star contamination is not heavy, and cluster sequences can be unambiguously detected, which is important for the extraction of radial profiles with small errors (see below). Using SBPs built with 2MASS images and a fit with \citet{King62} profile, \citet{Cohen07} derived the core radius in the \mbox{$\rm J$}\ band $\mbox{$\rm R_c$}(J)=61.5\arcsec\pm9.3\arcsec$. However, based on Hubble Space Telescope data and using a power-law plus core as fit function, \citet{NoGe06} derived $\mbox{$\rm R_c$}=3.7\arcsec$ in the equivalent V band, thus roughly resolving the post-core collapse nucleus. The post-core collapse state of NGC\,6397 does not affect the present analysis, since the goal here is the determination of changes produced in cluster radii derived under the assumption of a King-like profile (Sect.~\ref{Struc}) applied to RDP, MDP and SBPs built with different magnitude depths. We base the analysis of NGC\,6397 on \mbox{$\rm J$}, \mbox{$\rm H$}\ and \mbox{$\rm K_s$}\ 2MASS photometry extracted using VizieR\footnote{\em vizier.u-strasbg.fr/viz-bin/VizieR?-source=II/246} in a circular field of radius $\mbox{$\rm R_{ext}$}=70\arcmin$ centered on the coordinates provided in H03. This extraction radius is large enough to encompass the whole cluster, allowing as well for a significant comparison field. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig8.eps}} \caption{Structural analysis of NGC\,6397. Panel (a): decontaminated CMD of a central ($R<5\arcmin$) region. The reference magnitude $\mbox{$\rm J$}=15$ is indicated by the dashed-line. Shaded region: colour-magnitude filter. Background-subtracted RDPs for stars brighter than $\mbox{$\rm J$}<15+\Delta_{J15}$, with $\Delta_{J15}=1,~0,~-1,~-2$ are shown in panels (b) to (d), respectively. The respective King-like fits (solid-line) together with the fit uncertainty (shaded region) are shown.} \label{fig8} \end{figure} For a better definition of the cluster sequences we apply the statistical decontamination algorithm described in \cite{BB07}, which takes into account the relative number-densities of candidate cluster and field stars in small cubic CMD cells with axes corresponding to the magnitude \mbox{$\rm J$}\ and the colours \mbox{$\rm (J-H)$}\ and \mbox{$\rm (J-K_s)$}. Basically, the algorithm {\em (i)} divides the full range of magnitude and colours of the CMD into a 3D grid, {\em (ii)} computes the expected number-density of field stars in each cell based on the number of comparison field stars with magnitude and colours compatible with those in the cell, and {\em (iii)} subtracts the expected number of field stars from each cell. Typical cell dimensions are $\Delta\mbox{$\rm J$}=0.5$, and $\Delta\mbox{$\rm (J-H)$}=\Delta\mbox{$\rm (J-K_s)$}=0.25$, which are large enough to allow sufficient star-count statistics in individual cells and small enough to preserve the morphology of the CMD evolutionary sequences. The comparison field is the region located between $50\leq R(\arcmin)\leq70$, which is beyond the tidal radius. Field-decontaminated CMDs allow for a better definition of colour-magnitude filters, useful to remove stars (and artifacts) with colours compatible with those of the field which, in turn, improves the cluster/background contrast in RDPs and SBPs. They are wide enough to accommodate cluster MS and evolved star colour distributions and dynamical evolution-related effects, such as enhanced fractions of binaries and other multiple systems (e.g. \citealt{BB07}; \citealt{N188}). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{fig9.eps}} \caption{Left panels: RDP and SBP structural radii of NGC\,6397 as a function of $\Delta_{J15}$, normalised to the values measured in the deepest profile. Right panels: RDP to SBP radii ratios (similar to Fig.~\ref{fig5}).} \label{fig9} \end{figure} Figure~\ref{fig8} (panel a) displays the decontaminated CMD of a central region of NGC\,6397, with $R<5\arcmin$, somewhat larger than the half-light radius (Table~\ref{tab3}). We take $\mbox{$\rm J$}=15$ as reference to extract the depth-variable profiles. RDPs and SBPs are built with colour-magnitude filtered photometry, with the faint end varying in steps of $\Delta_{J15}=0.5$, with the deepest (i.e. at the available 2MASS depth) profile beginning at $\mbox{$\rm J$}=16$ and the brightest one ending near the giant clump at $\mbox{$\rm J$}=13$. The extracted profiles are fitted with the King-like function discussed in (Sect.~\ref{ModelSCs}). A selection of depth-limited RDPs, together with the respective fits and uncertainties, is shown in Fig.~\ref{fig8}, and the corresponding RDP and SBP (\mbox{$\rm J$}\ band) radii are given in Table~\ref{tab3}. Within uncertainties, the present value of the core radius (for the deepest profile), $\mbox{$\rm R_c$}(\mbox{$\rm J$})=1.4\arcmin\pm0.3\arcmin$, agrees with that derived by \citet{Cohen07}, using the same fit function. The near-infrared half-light radius, on the other hand, is larger than the optical one (H03), $\mbox{$\rm R_{hL}$}(\mbox{$\rm J$})\approx1.5\mbox{$\rm R_{hL}$}(V)$. Effects of the varying magnitude depth on the radii of NGC\,6397 are examined in Fig.~\ref{fig9}. Qualitatively, the resulting curves agree, within uncertainties, with the behaviour predicted by the GC models (Figs.~\ref{fig3} and \ref{fig5}). Compared to the values measured in the deepest RDP, the tidal (panel a), half-star counts (b) and core (c) radii decrease for shallower profiles, especially for $\Delta_{J15}\geq-0.5$, remaining almost uniform for $\Delta_{J15}<-0.5$. In particular, the core radius measured in shallow RDPs (containing essentially giants) drops to $\approx45\%$ of its deepest value (which includes stars at the top of the MS). Consistently with the GC models containing a spatially variable MF (Sect.~\ref{Struc}), the varying depth affects the tidal, half and core radii, with increasing intensity. SBP radii, on the other hand, remain essentially uniform with variable depth, consistent with the GC models (Sect.~\ref{Struc}). The same conclusions apply to the RDP to SBP radii ratio (right panels). \begin{table} \caption[]{Radii of NGC\,6397 from RDPs and 2MASS SBPs} \label{tab3} \renewcommand{\tabcolsep}{0.9mm} \renewcommand{\arraystretch}{1.25} \begin{tabular}{cccccccc} \hline\hline &\multicolumn{3}{c}{RDP}&&\multicolumn{3}{c}{SBP (\mbox{$\rm J$}\ band)}\\ \cline{2-4}\cline{6-8} $\Delta_{J15}$&\mbox{$\rm R_c$}&\mbox{$\rm R_{hSC}$}&\mbox{$\rm R_t$}&&\mbox{$\rm R_c$}&\mbox{$\rm R_{hL}$}&\mbox{$\rm R_t$} \\ (mag)&(\arcmin)&(\arcmin)&(\arcmin)&&(\arcmin)&(\arcmin)&(\arcmin)\\ (1)&(2)&(3)&(4)&&(5)&(6)&(7)\\ \hline $-2.0$&$1.3\pm0.1$&$3.8\pm0.1$&$33\pm5$&&$1.2\pm0.3$&$3.4\pm0.1$&$28\pm5$ \\ $-1.5$&$1.3\pm0.1$&$4.0\pm0.2$&$39\pm8$&&$1.2\pm0.3$&$3.4\pm0.1$&$30\pm5$ \\ $-1.0$&$1.3\pm0.1$&$4.0\pm0.2$&$42\pm8$&&$1.2\pm0.3$&$3.4\pm0.1$&$26\pm5$ \\ $-0.5$&$1.4\pm0.1$&$3.9\pm0.2$&$44\pm7$&&$1.2\pm0.3$&$3.4\pm0.1$&$27\pm8$ \\ $~0.0$&$1.7\pm0.1$&$4.0\pm0.1$&$41\pm4$&&$1.2\pm0.3$&$3.4\pm0.1$&$27\pm6$ \\ $+0.5$&$2.3\pm0.1$&$4.4\pm0.1$&$40\pm4$&&$1.4\pm0.4$&$3.4\pm0.1$&$28\pm8$ \\ $+1.0$&$2.9\pm0.1$&$4.9\pm0.1$&$48\pm3$&&$1.4\pm0.3$&$3.5\pm0.1$&$32\pm2$ \\ \hline \end{tabular} \begin{list}{Table Notes.} \item Core and tidal radii were derived from fits of \citet{King62} functions (Sect.~\ref{Struc}) to the respective profiles. The half-star counts and half-light radii were measured directly on the profiles. \end{list} \end{table} \section{Concluding remarks} \label{Conclu} In this work we simulated star clusters of different ages, structure and mass functions, assuming that the spatial distribution of stars follows an analytical function, similar to \citet{King62} profile. The mass and near-infrared luminosities of each star were assigned according to a mass function with a slope that may depend on distance to cluster center. They form the set of models from which we built number-density, mass-density and surface-brightness profiles, allowing for a variable photometric depth. The structural parameters core, half-light, half-mass and half-star count, and tidal radii, together with the concentration parameter, were measured in the resulting radial profiles. Next we examined relations among similar parameters measured in different profiles, and determined how each parameter depends on photometric depth. We point out that the results should be taken as upper-limits, especially for open clusters, since we have considered noise-free photometry and a large number of stars, which produced small statistical uncertainties. With respect to the adopted form of the radial distribution of stars, we note that \citet{King62} isothermal sphere, single-mass profile has been superseded by more realistic models like those of \citet{King66}, \citet{Wilson75} and \citet{EFF87}, which have been fit mostly to the SBPs of Galactic and extra-Galactic GCs (Sect.~\ref{ModelSCs}). The analytical functions associated with these models are characterised by different scale radii (among other parameters) that are roughly related to \citet{King62} radii. Thus, it is natural to extend the scaling with photometric depth undergone by \citet{King62} radii to the equivalent ones in the other models. The main results can be summarised as follows. \begin{itemize} \item {\em (i)} Structural parameters derived from surface-brightness profiles are essentially insensitive to photometric depth, except perhaps the cluster radius in very young clusters. \item {\em (ii)} Uniform mass functions also result in structural parameters insensitive to photometric depth. \item {\em (iii)} Number-density and mass-density profiles built with shallow photometry result in underestimated radii, with respect to the values obtained with deep photometry. Tidal, half-star count and half-mass, together with the core radii are affected with increasing intensity. \item {\em (iv)} Because of the presence of bright stars, radii underestimation increases for young ages. \item {\em (v)} For clusters older than $\sim1$\,Gyr, number-density and mass-density radii present essentially the same values; for younger ages, RDP radii become increasingly larger than MDP ones, especially at the deepest profiles. \item {\em (vi)} Irrespective of age, profiles deeper than the turnoff have RDP radii systematically larger than SBP ones, especially the core. \item {\em (vii)} The concentration parameter also changes with photometric depth, reaching a maximum around the turnoff. \end{itemize} Most of the above model predictions were qualitatively confirmed with radii measured in ground-based RDPs and SBPs of the nearby GC NGC\,6397. In principle, working with SBPs has the advantage of producing more uniform structural parameters, since they are almost insensitive to photometric depth. However, as discussed in Sect.~\ref{intro}, SBPs usually present high levels of noise at large radii. Noise that is also present in SBPs of clusters projected against dense fields and/or the less populous ones. A natural extension of this work would be to examine radial profiles built with photometry that includes observational uncertainties, differential absorption, metallicity gradients, binaries, and star cluster models with a number of stars compatible with those of open clusters. As a consequence of the wide range of distances to the Galactic (and especially extra-Galactic) star clusters, interstellar absorption, and intrinsic instrumental limitations, the available photometric data for most clusters do not sample the low-mass stars. All sky surveys like 2MASS, usually are restricted to the giant branch, or the upper main sequence, for clusters more distant than a few kpc. In such cases, the structural parameters have to be derived from radial profiles built with photometry that does not reach low-mass stars. The present work provides a quantitative way to estimate the intrinsic (i.e. in the case of photometry including the lower main sequence) values of structural radii of star clusters observed with depth-limited photometry. \begin{acknowledgements} We thank the anonymous referee for helpful suggestions. We acknowledge partial support from the Brazilian institution CNPq . \end{acknowledgements}
{ "timestamp": "2007-11-19T13:54:32", "yymm": "0711", "arxiv_id": "0711.2919", "language": "en", "url": "https://arxiv.org/abs/0711.2919" }
\section{Introduction} Accretion disks can carry small- and large-scale magnetic fields. The small-scale field ($\ell\la R$, where $\ell$ is the field scale length and $R$ measures the radial distance from the disk center) can be locally generated by the MHD dynamo (Brandenburg et al. 1995; Stone et al. 1996) supported by the turbulence, which results from the magneto-rotational instability (MRI, Balbus \& Hawley 1991). This field can provide the outward transport of angular momentum in the bulk of the disk with the help of local Maxwell stresses (Shakura \& Sunyaev 1973; Hawley, Gammie, \& Balbus 1996). The large-scale field ($\ell > R$) is unlikely produced in accretion disks (however, see Tout \& Pringle 1996), and can either be captured from the environment and dragged inward by an accretion flow (Bisnovatyi-Kogan \& Ruzmaikin 1974, 1976), or inherited from the past evolution (see \S 4). The large-scale field can remove the angular momentum from accretion disks by global Maxwell stresses through the magnetized disk corona (K\"onigl 1989). A large-scale bipolar field, unlike a small-scale field, can not dissipate locally due to the magnetic diffusivity and can not be absorbed by the central black hole. In the case of inefficient outward diffusion of the bipolar field through the disk (see Narayan, Igumenshchev, \& Abramowicz 2003; Spruit \& Uzdensky 2005; Bisnovatyi-Kogan \& Lovelace 2007; and for other possibility, see van Ballegooijen 1989; Lubow, Papaloizou, \& Pringle 1994; Lovelace, Romanova, \& Newman 1994; Heyvaerts, Priest, \& Bardou 1996; Agapitou \& Papaloizou 1996; Livio, Ogilvie, \& Pringle 1999), this field is accumulated in the innermost region of an accretion disk and forms a ``magnetically arrested disk,'' or MAD (Narayan et al. 2003). The MAD consists of two parts: the outer, almost axisymmetric, Keplerian accretion disk and the inner magnetically dominated region, in which the accumulated vertical field disrupts the outer disk at the magnetospheric radius $R_{\rm m}\sim 8\pi GM\rho/B^2$, where $M$ is the central mass, $\rho$ is the accretion mass density, and $B$ is the magnetic induction. It is believed that the large-scale bipolar field in accretion disks is responsible for the formation of jets observed in a large variety of astrophysical objects (e.g., Livio, Pringle, \& King 2003). The magnetically driven jets can be of two types, basically depending on a mass load by the disk matter (e.g., Lovelace, Gandhi, \& Romanova 2005): Poynting jets and hydromagnetic jets, which have, respectively, small and large mass loads. The hydromagnetic jets can be formed by two mechanisms: the magneto-centrifugal mechanism (Blandford \& Payne 1982; K\"onigl \& Pudritz 2000) and the toroidal-field pressure generated by the disk rotation (Lynden-Bell 2003; Kato, Mineshige, \& Shibata 2004). The magneto-centrifugal mechanism produces relatively wide outflows and, to be consistent with observations of the collimated jets, requires an additional focusing mechanism. Jets driven by the toroidal-field pressure can have a high degree of collimation, but these jets are known to be kink unstable (Eichler 1993; Appl 1996; Spruit, Foglizzo, \& Stehle 1997). Poynting jets are naturally self-collimated and expected to be marginally kink stable (Li 2000; Tomimatsu, Matsuoka, \& Takahashi 2001). These jets can originate in the innermost region of accretion disks and powered either by the disks themselves (Lovelace, Wang, \& Sulkanen 1987; Lovelace et al. 2002) or by rotating black holes (Blandford \& Znajek 1977; Punsly 2001; also see, Takahashi et al. 1990; Komissarov 2005; Hawley \& Krolik 2006; McKinney 2006). Our study is based on two- and three-dimensional (2D and 3D, respectively) MHD simulations and has two main goals. First, we investigate the dynamics and structure of MAD's. We show that MAD's are formed in the accretion flows, which carry inward large-scale poloidal magnetic fields. Inside the magnetospheric radius $R_{\rm m}$, the matter accretes as discrete streams and blobs, fighting its way through the strong vertical magnetic field fragmented in separate bundles. Because of rotation, the streams take spiral shapes. Second, we demonstrate a link between the existence of MAD's and production of powerful Poynting jets. These jets should always be generated in MAD's because of the interaction of the spiraling accretion flow with the vertical magnetic bundles, which, as the result, are twisted around the axis of rotation. This paper extends the work of Igumenshchev, Narayan, \& Abramowicz (2003) by studying in more detail the radiatively inefficient accretion disks with poloidal magnetic fields. We employ a new version of our 3D MHD code, which can be utilized in multi-processor simulations. The paper is organized as follows: We describe the solved equations, the numerical method used, and initial and boundary conditions in \S 2. We present our numerical results in \S 3, and discuss and summarize them in \S 4. \section{Numerical method} We simulate nonradiative accretion flows around a Schwarzschild black hole of mass $M$ using the following equations of ideal MHD: \begin{equation} {d\rho\over dt} + \rho{\bf\nabla\cdot v} = 0, \end{equation} \begin{equation} \rho{d{\bf v}\over dt} = -{\bf\nabla} P - \rho{\bf\nabla}\Phi + {1\over 4\pi}({\bf\nabla}\times{\bf B})\times {\bf B}, \end{equation} \begin{equation} {\partial\over\partial t}\left(\rho{v^2\over 2}+\rho\epsilon+ {B^2\over 8\pi}+\Phi\right)=-\nabla\cdot{\bf q}, \end{equation} \begin{equation} {\partial{\bf B}\over \partial t} = {\bf\nabla}\times({\bf v}\times{\bf B}), \end{equation} where ${\bf v}$ is the velocity, $P$ is the gas pressure, $\Phi$ is the gravitational potential, $\epsilon$ is the specific internal energy of gas, and ${\bf q}$ is the total energy flux per unit square (see, e.g., Landau \& Lifshitz 1987). We adopt the ideal gas equation of state, \begin{equation} P=(\gamma-1)\rho\epsilon, \end{equation} with an adiabatic index $\gamma=5/3$. We neglect self-gravity of the gas and employ a pseudo-Newtonian approximation (Paczy\'nski \& Wiita 1980) for the black hole potential \begin{equation} \Phi=-{GM\over R-R_g}, \end{equation} where $R_g=2GM/c^2$ is the gravitational radius of the black hole. No explicit resistivity and viscosity are applied in equations (2)-(4). However, because of the use of the total energy equation (3), the energy released due to the numerical resistivity and viscosity is consistently accounted as heat in our simulations. The MHD equations (1)-(4) are solved employing the time-explicit Eulerian finite-difference method, which is an extension to MHD of the hydrodynamic piecewise-parabolic method by Colella \& Woodward (1984). We solve the induction equation (4) using the constrained transport (Evans \& Hawley 1988; Gardiner \& Stone 2005), which preserves the $\nabla\cdot{\bf B}=0$ condition. In our method, we employ the approximate MHD Riemann solver by Li (2005). Test simulations have shown that this solver is robust and provides a good material interface tracking. We use the spherical coordinates $(R,\theta,\phi)$. Our 3D numerical grid has $182\times 84\times 240$ zones in the radial, polar, and azimuthal directions, respectively. The radial zones are spaced logarithmically from $R_{in}=2 R_{g}$ to $R_{out}=220 R_{g}$. Both hemispheres are considered, in which polar cones with the opening angle $\pi/8$ are excluded. Therefore, the polar domain extends from $\theta=\pi/16$ to $15\pi/16$. The grid resolution in the polar direction is gradually changed from a fine resolution around the equatorial plane to a coarse resolution near the poles (with the maximum-to-minimum grid-size ratio $\approx 3$). The azimuthal zones are uniform and cover the full $2\pi$ range in $\phi$. The absorption condition for the mass and the condition of the zero-transverse magnetic field are applied in the inner and outer radial boundaries, providing that the mass and field can freely leave the computational domain through these boundaries. In the boundaries around the excluded polar regions, we apply the reflection boundary conditions, which means that no streamlines and magnetic lines can go through these boundaries. At the beginning of our simulations, the computational domain is filled with a very low-density, nonmagnetized material. The simulations are started in 2D, assuming the axial symmetry, with an injection of mass in a slender torus, which is located in the equatorial plane at $R_{inj}=210\, R_{g}$; i.e., close enough to $R_{out}$. This mass has the Keplerian angular momentum and specific internal energy $\epsilon_{inj}=0.045\,GM/R_{inj}$. After an initial period of simulation without magnetic fields, the mass forms a steady thick torus, which has the inner edge at $R\approx 150\,R_{g}$ and which outer half is truncated at $R_{out}$. The torus contains a constant amount of mass and is in a dynamic equilibrium: all the injected mass flows outward through $R_{\rm out}$ after a circulation inside the torus. No accretion flow at this point is formed. This hydrodynamic, steady, thick torus is used as an initial configuration in our MHD simulations. The MHD simulations are started at $t=0$ from the steady, thick torus by initiating the injection of a poloidal magnetic field into the injection slender torus at $R_{inj}$. The numerical procedure for the field injection is similar to that described by Igumenshchev et al. (2003) except for one modification: now the strength of the injected field can be limited by setting the minimum $\beta_{\rm inj}$, which is the ratio of the gas pressure to the magnetic pressure at $R_{\rm inj}$. This modification allows us to better control the rate of field injection. The entire volume of the thick torus is filled by the field during about one orbital period, $t_{\rm orb}$, estimated at $R_{\rm inj}$. Since this moment, $t\simeq t_{orb}$, the formation of accretion flow begins as a result of redistribution of the angular momentum in the torus due to Maxwell stresses. In the following discussion, we will use the time normalized by the time-scale $t_{orb}$, i.e. $t\rightarrow t/t_{\rm orb}$. \section{Results} We present the results of combined axisymmetric 2D and non-axisymmetric 3D simulations. The initial evolution in our models has been simulated in 2D. This allows us to consider longer evolution times in comparison to those that can be obtained in 3D simulations, because of the larger requirements for computational resources in the latter case. We initiate 3D simulations starting from developed axisymmetric models. The results have shown that non-axisymmetric motions are not very important in the outer parts of the constructed accretion flows and, therefore, the use of 2D simulations on the initial evolution stages is the reasonable simplification. We consider three models, which differ by the rates of field injection determined by $\beta_{\rm inj}=10$, 100, and 1000, and we will refer to these models as Model~A, B, and C, respectively. All other properties of the models, including the injection radius $R_{inj}$, internal energy $\epsilon_{inj}$, and numerical resolution, are the same (see \S 2). \subsection{Accretion flows} The initial axisymmetric development of the models is qualitatively similar: the inner edge of the thick torus is extended toward the black hole, forming relatively thin, almost Keplerian accretion disks. The time of the disk development is varied, depending on the strength of the injected field. The accretion of the mass into the black hole begins at $t\approx 0.7$ in Model~A, $\approx 1.3$ in Model~B, and $\approx 4.2$ in Model~C (time is given in units of the orbital period at $R_{inj}$, see \S 2). At this stage, the evolution of the disks is governed mainly by global Maxwell stresses produced by the poloidal field component. This component is advected inward with the accretion flow and, because of the disks' Keplerian rotation, generates relatively strong toroidal magnetic fields localized above and below the mid-plane. These toroidal fields form a highly magnetized disk corona with a typical $\beta\sim 0.01$. Model~B and, especially Model~C, demonstrate the development of 2D MRI. This development is similar to that observed by Stone \& Pringle (2001); in particular, in their ``Run C." We have found the origin of the channel solution (see Hawley \& Balbus 1992) in the central regions of Models~B and C. This solution consists of oppositely directed radial streams and is the characteristic feature of the axisymmetric non-linear MRI (Stone \& Norman 1994). Analysis of the models shows that the channel solution is developed when the wavelength $\lambda=2\pi V_A/\sqrt{3}\Omega$ of the fastest growing mode of the MRI is well resolved on the numerical grid, i.e. $\lambda\ga 5\Delta x$, where $\Delta x$ is the grid size, $V_A$ is the Alfv\'en velocity, and $\Omega$ is the angular velocity. Model~A shows no indication of the MRI, which can be attributed to the strong magnetic fields, which suppress the instability. In this model, the estimate of $\lambda$ typically exceeds the disk thickness. Model~A has some resemblance to nonturbulent ``Run F'' of Stone \& Pringle (2001). In spite of the mentioned similarities with the results of Stone \& Pringle (2001), our models show different behavior on the long evolution times. Our simulation design with the permanent injection of mass and magnetic field results in accretion disks, which accumulate the poloidal field in the center and form MAD's. The models of Stone \& Pringle (2001; also De Villiers, Hawley, \& Krolik 2003; Hirose et al. 2004; McKinney \& Gammie 2004; Hawley \& Krolik 2006) did not form MAD's and did not show a long-time accretion history, probably because of the initiation of simulations from static magnetized tori, which contain a limited amount of mass and magnetic flux of one sign. We will concentrate on the results describing the formation, evolution, and structure of MAD's in the following text. Other aspects of our results will be reported elsewhere. Figure~1 shows example snapshots of the axisymmetric density distribution in Model~B from 2D simulations at two successive moments, $t=5.1153$ and $5.1458$. The accretion flow is nonuniform because of the development of turbulence. The turbulence results from the combined effect of the MRI and current sheet instability. The latter instability locally releases heat due to reconnections of the oppositely directed toroidal magnetic fields. The reconnection heat makes a significant contribution to the local energy balance in the central regions of the flow, because of the relatively high energy density of the field, which is comparable to the gravitational energy density of the accretion mass. The thick disk structure observed in Fig.~1 is explained by convection motions supported by the reconnection heat. Note that the case, in which the turbulence is supported by only convection motions from the reconnections, without the effects of rotation and MRI, had been demonstrated in simulations of spherical magnetized accretion flows (Igumenshchev 2006). In the case of disk accretion, almost axisymmetric convection motions, similar to those found here, had been observed in 3D models with toroidal magnetic fields (Igumenshchev et al. 2003). The convection motions in Models~B and C make these models relevant to convection-dominated accretion flows (Narayan, Igumenshchev, \& Abramowicz 2000; Quataert \& Gruzinov 2000). Our simulations show that the poloidal field is transported inward in axisymmetric turbulent flows and accumulated in the vicinity of the black holes. When the central poloidal field reaches some certain strength (about equipartition with the gravitational energy of the accreting mass), the accretion flow becomes unstable (Narayan et al. 2003). In axisymmetric simulations, the instability takes the form of cycle accretion, in which the more-extended periods of halted accretion (see Fig.~1a) are followed by the relatively short periods of accretion (see Fig.~1b). In the case of the halted accretion period, the inner accretion disk is truncated at the magnetospheric radius $R_{m}$, which is $\approx 15\,R_g$ in Fig.~1a. The pressure of the strong central vertical field (see Fig.~2a) prevents the mass accumulated at $R_{m}$ from falling into the black hole. The accretion begins as soon as the gravity of the accumulated mass overcomes the magnetic pressure. During the accretion period, the whole magnetic flux, which is localized inside $R_{\rm m}$ in the period of halted accretion, is moved on the black hole horizon (see Fig.~2b). Note that similar structural features of the inner MHD flows in accretion disks related to the model of gamma-ray bursts were discussed by Proga \& Zhang (2006). Figure~3 illustrates the time dependence of the accretion flow in Model~B, showing the evolution of the mass accretion rate $\dot{M}_{in}$ and magnetic fluxes in the midplane inside the five specific radii: $210\,R_g$ ($=R_{inj}$), $100\, R_g$, $50\, R_g$, $25\, R_g$, and $2\, R_g$ ($=R_{inj}$). This figure shows the evolution, which has been simulated in 2D from $t=0$ to $2.14$ and in 3D after $t=2.14$. The vertical dashed line in Fig.~3 indicates the moment of initiation of the 3D simulations. The cycle accretion in the 2D simulations begins at $t\approx 1.4$ and is clearly seen as a sequence of spikes in the time-dependence of $\dot{M}_{in}$ (see Fig.~3a). Spikes, which are related to the same cycle accretion, are also observed in the variation of magnetic flux inside $R=2\, R_g$ (see Fig.~3b). The magnetic fluxes inside the other selected radii are gradually increased with time because of the inward advection of the vertical field. The time dependence of these fluxes is not significantly influenced by the cycle accretion. The structure and dynamics of the inner region in Model~B are drastically changed in the 3D simulations. Shortly after the initiation of the 3D simulations at $t=2.14$, the axisymmetric distribution of mass near $R_{\rm m}$ undergoes the Rayleigh-Taylor and, possibly, Kelvin-Helmholtz instability (see Kaisig, Tajima, \& Lovelace 1992; Spruit, Stehle, \& Papaloizou 1995; Chandran 2001; Li \& Narayan 2004) with the fastest growing azimuthal mode number $m\simeq 50$ (the latter is probably determined by our grid resolution). As a result, the empty region inside $R_{\rm m}$ is quickly filled, on the free-fall time scale estimated at $R_{\rm m}$, with the large number of density spikes moved almost radially toward the center. These spikes quickly disappear in the black hole and, at a later time, the inner disk structure is modified toward establishing a dynamic quasi-steady state. This state is characterized by a low $m$-mode ($m\simeq 1$-5) spiral-flow structure, which results from the interaction of the accreting mass with the strong vertical magnetic field. Note that the similar low $m$-mode flow structure was found in the simulations of accretion flows onto a magnetic dipole (Romanova \& Lovelace 2006). The non-axisymmetric inner flow is highly time-variable and experiences a quasi-periodic behavior. Figure~4 shows an example of the flow structure inside 50 $R_g$ in Model~B, at two successive moments: $t=2.2767$ and 2.2867. The flow is essentially 3D inside the magnetically dominated region limited by the radius $\simeq 35\,R_g$ and remains almost axisymmetric on the outside of this radius. In the magnetically dominated region, the flow forms moderately tightened spirals of dense matter, which are clearly seen in Figs~4a and 4b. This matter is quickly accreted into the black hole with the radial velocity, which is $\sim 0.5$ a fraction of the free-fall velocity. Such a relatively fast infall is explained by the efficient loss of the angular momentum by the mass during its interaction with the vertical field. The field is distributed nonuniformly in the disk plane, concentrating in bundles that penetrate through the plane in very low density, magnetically dominated (with $\beta\sim 0.01$) regions, or magnetic ``islands.'' The rotating mass interacts with magnetic bundles and forces them to twist around the disk's rotational axis. In the simulations, this twist is observed as the rotation of magnetic islands around the center in the disk plane. The rotational velocity of the islands typically has the reduced rotational velocity by the factor of $\sim0.5-1$ in comparison with the velocity of the surrounding accretion matter. This can be explained by the resistance of the large-scale vertical field to such a twist. The faster rotation of accretion matter and slower rotation of magnetic islands produces a shear flow. The shear flow plays two roles in our simulations. First, it provides the exchange of momentum and energy between the accreting mass and vertical field. Second, the shear flow results in an ablation of the islands caused by magnetic diffusivity, making each island a temporal structure. An example evolution of magnetic islands can be seen in Fig.~4: the magnetic islands observed as low-density spiral arms above the center in Fig.~4a are observed below the center in Fig.~4b, after about half a revolution in the clockwise direction. In the latter figure, the islands are apparently reduced in size due to the ablation. The vertical field ablated from magnetic islands is carried inward by the accretion flow and accumulates on the black hole horizon. This accumulation results in quasi-periodic eruptions of the field outward from the horizon as soon as the field pressure overcomes the dynamic pressure of the accreting mass. The eruptions typically take the form of high-velocity narrow streams (in the equatorial cross-section) of a low-density, magnetically dominated medium fountained outward from the black hole. In Fig.~4b, four magnetic islands observed as low-density regions inside $R\approx 15\, R_g$ result from such eruptions and the eruption of one of these islands (to the right from the center; see also the steam that produced it) still continues at the moment shown. In the consequent evolution, these islands are pushed outward and stretched in the azimuthal direction by the accretion flow, and take the spiral shape similar to that shown in Fig.~4a. Model~A evolves faster and accumulates a larger magnetic flux at the center in comparison with Model~B. Figure~5 shows the evolution of the accretion rate $\dot{M}_{\rm in}$ and magnetic fluxes in Model~A. Qualitatively, the evolution of these quantities is similar to the evolution of those in Model~B (see Fig.~3). Quantitatively, however, Model~A demonstrates significantly larger time-averaged accretion rates (by about two orders of magnitude) and longer quasi-periods of the cycle accretion (represented by the intervals between spikes in the time-dependence of $\dot{M}_{\rm in}$ in Fig.~5a) in the 2D simulations. By the end of the 2D simulations at $t\approx 1.7$, this model has the maximum $R_m\simeq 30-40\,R_g$. In the 3D simulations, Model~A experiences the initial transient period, similar to the period of the development of the Rayleigh-Taylor instability in Model~B (see above), in which the non-axisymmetric, small-scale structures quickly appear and disappear. Figure~6 shows an example of the developed low $m$-mode spiral structure in Model~A obtained after the transient period. This structure is clearly dominated by the $m=1$ mode. The magnetically dominated region is extended up to $R\simeq 70\, R_g$. Note that the spiral-density arms seen in Fig.~6 are more open than the arms in Model~B (see Fig.~4a). This could be due to the stronger central field in Model~A. Model~C is our slowest evolving model and, accordingly, shows the slowest rate of accumulation of the central vertical field. This model has been calculated only in 2D and demonstrated the qualitative similarity to the axisymmetric evolution of Models~A and B. The cycle accretion, which is caused by the accumulated field, begins at $t\approx 4.8$ in Model~C. The model demonstrates more-efficient turbulent motions in the accretion flow. This can be attributed to weaker magnetic fields, which suppress less the MRI and convection motions. At the end of simulation at $t\approx 6$, the model has the maximum $R_m\simeq 6\, R_g$. \subsection{Poynting Jets} The 3D simulations of Models~A and B show that the vertical field penetrated the central magnetically dominated regions in MAD's is twisted around the axis of rotation by the rotating accretion flows. The field twist generates electromagnetic perturbations, which propagate outward and transport the released energy in the form of a Poynting flux (e.g., Landau \& Lifshitz 1987) \begin{equation} {\bf S}={1\over 8\pi}(({\bf v}\times{\bf B})\times{\bf B}). \end{equation} The Poynting flux is distributed nonuniformly in the polar angles, basically showing two components: a jet-like concentration of the flux near the poles and a wide-spread distribution of the flux in the equatorial and mid-polar-angle directions (from $\theta\sim\pi/4$ to $\sim 3\pi/4$). Figures~7 and 8 show example $\theta$-distributions of the radial Poynting flux (solid lines), \begin{equation} S_R=v_R{B^2\over 4\pi}-{B_R\over 4\pi}({\bf v}\cdot {\bf B}), \end{equation} at six different radii, 5 $R_g$, 10 $R_g$, 25 $R_g$, 50 $R_g$, 100 $R_g$, and 220 $R_g$, which cover most of the radial domain in Models~B and A, respectively. The shown distributions are averaged in the azimuthal direction and in time over the interval $\Delta t\simeq 0.05$, using a set of data files stored during the simulations. Outside of the inner magnetically dominated region (at $R\ga 35\,R_g$ in Model~B and $R\ga 70\,R_g$ in Model~A), the outward (positive) Poynting flux in the equatorial and mid-polar-angle directions is supported mostly by the rotation of the outer, almost axisymmetric disks, in which the poloidal field component is frozen in. Such a flux is present in both the axisymmetric 2D and 3D simulations. The 3D simulations introduce new important features in the Poynting flux distribution: an increase of the flux at the equatorial and mid-polar-angle directions inside the magnetically dominated region, and at the polar directions outside this region (see Figs~7 and 8). The equatorial flux inside the magnetically dominated region is generated due to the twist of the vertical field by the spiraling non-axisymmetric accretion flows. In Model~B, this flux, represented by bumps in the $\theta$-distributions, gradually deviates from the equatorial plane toward the poles as the radial distance increases (see Figs~7a-7d). At $R\ga 100\,R_g$, the flux is collimated into bi-polar Poynting jets (see Figs~7e and 7f). In the case of Model~A, the process of jet collimation is less evident and somewhat different; but still, one can observe the formation of narrow bi-polar Poynting jets starting from $R\simeq 10\,R_g$ and further development of these jets at large radii (see Figs~8b-8f). Figures~7 and 8 show $\theta$-distributions of the kinetic flux (dashed lines), \begin{equation} F_R=v_R\rho{v^2\over 2}, \end{equation} for comparison with the Poynting flux. Typically, the kinetic flux is comparable, but does not exceed the Poynting flux in the polar jets (except in the outermost region in Model~A, see Fig.8f). Accordingly, the jet velocity is mostly sub-Alfv\'enic. However, the value of the kinetic flux is relatively large and this is in some contradiction with our expectations that MAD's can develop Poynting flux dominated jets. The problem of the excessive kinetic flux in our simulations can probably be explained by the action of the numerical magnetic diffusivity (see \S 4), which results in an unphysically large mass load of the Poynting jets and the consequent excessive kinetic flux in them. The Poynting jets are powered by the released binding energy of the accretion mass. To estimate quantitatively the amount of energy going into the jets, we calculate the Poynting jet ``luminosity" $\dot{E}_{jet}$ as a function of the radius R, \begin{equation} \dot{E}_{jet}(R)=\int R^2 S_R\,d\Omega, \end{equation} where the integration is taken over the solid angles $\Omega$ occupying the polar regions with $\theta < \pi/4$ and $\theta > 3\pi/4$ (excluding the boundary polar cones, see \S 2). For comparison, we also calculate the Poynting total luminosity $\dot{E}_{tot}$, which is defined analogously to $\dot{E}_{jet}$, but with the integration in eq.~(9) taken over the whole sphere. Figures~9 and 10 show the radial profiles of the normalized $\dot{E}_{jet}$ (solid lines) and $\dot{E}_{tot}$ (dashed lines) in Models~B and A, respectively. The jet luminosity $\dot{E}_{jet}$ weakly depends on the radius at $R\ga 50\,R_g$ in both models and equals to $\approx 1.5\%$ in Model~B and $\approx 0.5\%$ in Model~A. Here, we quantify the luminosity in the units of accretion power $\dot{M}_{in}c^2$. The total luminosity $\dot{E}_{tot}$ includes the flux from the bi-polar Poynting jets and wide equatorial Poynting outflow. The latter component of $\dot{E}_{tot}$ exceeds $\dot{E}_{jet}$ by the factor of $\sim 3$ at large radii (see Figs~9 and 10). This can be the consequence of the employed simulation design (see \S 2), in which the disk accretion at outer radii is mostly provided by the global Maxwell stresses. The smaller value of the final (at large radii) relative $\dot{E}_{jet}$ in Model~A, in comparison with that in Model~B, can be explained by the different structure of the inner magnetically dominated region in these models. Model~A has the less tighten spiral density arms (see Fig.~6), in which the mass accretes with larger radial velocity and, therefore, delivers less energy to the field. Note, also, that the value of $\dot{E}_{jet}$ in Model~A takes the relatively large finite value right at the inner boundary $R_{in}$ (see Fig.~10), whereas $\dot{E}_{jet}$ in Model~B begins from a small value at $R_{in}$ and gradually increases outward (see Fig.~9). This difference in the behavior of $\dot{E}_{jet}$ can be attributed to the discussed difference of the innermost structure in the considered models. \section{Discussion and Conclusions} We have performed a numerical study of the formation and evolution of quasi-stationary MAD, which is characterized by a strong vertical magnetic field accumulated at the disk center. We employ the simulation design, in which the poloidal magnetic field of one sign is permanently injected into the computational domain at the outer numerical boundary and the unipolar vertical field is transported inward by the accretion flow. The accumulated field has a significant impact on the inner flow structure and dynamics in both 2D and 3D simulations. In the axisymmetric 2D simulations, the field pressure can temporarily halt the mass falling into the black hole, resulting in the cycle accretion, in which the longer periods of accumulation of the mass at the magnetospheric radius $R_m$ are followed by the short periods of accretion. The 3D simulations have shown, however, that the axisymmetric cycle accretion is not realized. Instead, the accumulated field causes the mass to accrete quasi-regularly in the form of non-axisymmetric spiral streams and blobs. We have demonstrated that 3D MAD's can be efficient sources of collimated, bipolar Poynting jets, which originate in the vicinity of the central black hole. These jets develop due to and are powered by the interaction of the spiral mass inflows with the central field split into separate magnetic bundles. The efficiency of conversion of the accretion energy $\dot{M}_{in} c^2$ into the Poynting jet energy is up to 1.5\% in our simulations. This estimate may not be accurate (we believe, underestimated) because of the use of the pseudo-Newtonian approximation (see \S 2). The better estimate of the efficiency can be obtained using general relativistic MHD simulations. We have presented the simulation results from three models of radiatively inefficient accretion disks, which differ by the strength of the injected field. In accordance with the previous studies (e.g., Stone \& Pringle 2001), the structure of the outer disks in these models is determined by the field strength. In Model~A, which has the largest injected field, the MRI is suppressed and the accretion flow is driven by global Maxwell stresses, which transport the excessive angular momentum outward from the disk through the disk corona. In Models~B and C, which have the smaller injected fields, the MRI and turbulent motions are developed. The turbulence in these models is axisymmetric and partially supported by the efficient convection motions resulting from dissipation of toroidal and small-scale, poloidal magnetic fields. These motions cause the increase of the disk thicknesses in comparison with non-turbulent Model~A. The 3D simulations of Models~A and B have demonstrated that non-axisymmetric motions are not important in the outer parts of the disks, outside the inner magnetically dominated region (the disks remain almost axisymmetric), but very important inside this region, resulting in the development of the spiral accretion flows and bi-polar Poynting jets. Numerical magnetic dissipations and reconnections result in a magnetic diffusivity, which influences the structure and dynamics of our models. The spatial scale, on which the diffusivity occurs in our simulations, is determined by the gridsize, which greatly exceeds the scales of various resistive mechanisms (e.g., Coulomb collisions, dissipative plasma instabilities) in the relevant astrophysical conditions. Therefore, the magnetic diffusivity is significantly overestimated in our models and results in an excessive slippage of an accretion flow through magnetic field. This slippage reduces the ability of the flow to drug inward the vertical field, but, however, the numerical diffusivity is not efficient enough to totally prevent the field accumulation at the disk center. The numerical diffusivity suppresses the MRI on the scales of the gridsize, and, therefore, prevents the development of turbulence in the outer regions of our models, where the gridsize is increased. Other effect of the numerical magnetic diffusivity is the enhancement of the ablation of magnetic islands, which are found in the 3D simulations (see \S 3.1). To test the sensitivity of our models to magnetic dissipations, we have performed a 2D simulation of the model, which is similar to Model~A, but has the double number of grid points in the $R$- and $\theta$-directions. The simulation has demonstrated the qualitative similarity of axisymmetric evolution of the high resolution model and Model~A: both models show the formation of accretion disks, accumulation of the vertical field in the disk centers, and development of the cycle accretion. The high resolution model forms a thiner laminar accretion disk. Unfortunately, a more detailed quantitative comparison of these models meets some difficulties because of the different properties of the mass and field injection region (see \S 2), which are changed with the change of the resolution. The main results of our study, the formation of MAD's and Poynting jets, have been obtained under the assumption of radiatively inefficient flows, but, we believe that these results can also be applied to the radiatively efficient, dense accretion disks (e.g., Kato, Fukue, \& Mineshige 1998). The formation of MAD's should not be affected by the radiative cooling as soon as the central field satisfies the equipartition condition, \begin{equation} {B^2\over 8\pi} \sim {GM\rho\over R_g}, \end{equation} where $\rho$ is the mass density in the innermost region. The radiative losses results in the higher $\rho$ and, therefore, the larger B is necessary to obtain radiative MAD's. We expect that the qualitatively similar spiral structure of the inner magnetically dominated region, to that found in our simulations, can be developed in the case of the radiative disks. Poynting jets should be a necessary attribute of the radiative MAD's as well. The formation of MAD's in the radiative disks can be used to explain the observations of the low/hard state in black hole binaries (for a review, see Remillard \& McClintock 2006). Here, we briefly discuss basic moments of this application of MAD's and leave more quantitative considerations for future works. We assume, for example, the development of the MAD in the radiation pressure dominated accretion disk at the subcritical regime (see Shakura \& Sunyaev 1973). In such a disk, the radiation diffusion time scale $t_{rad}\simeq H^2\sigma_T\rho/c$ can significantly exceed the Keplerian time $t_{K}= 2\pi R^{3/2}/\sqrt{GM}$ at small values of the $\alpha$-parameter, $\alpha \la 0.1$, in virtu of the relation \begin{equation} {t_{K}\over t_{rad}} \simeq 3.4\alpha, \end{equation} which follows from the Shakura-Sunyaev solution. Here, we denote $H$ to be the disk half-thickness and $\sigma_T=0.4$ cm$^2$/g to be the Thomson scattering cross-section. As soon as $t_{rad}\gg t_{K}$ in the outer Shakura-Sunyaev disk, $t_{rad}$ can also significantly exceed the accretion velocity $t_{accr}\sim t_{K}$ in the inner spiral flow in the MAD, in virtu of the relation \begin{equation} {t_{accr}\over t_{rad}}\propto R, \end{equation} which is satisfied in accretion flows with the scaling law of the accretion velocity $v_{accr}\propto R^{-1/2}$ and $H\propto R$. Having $t_{rad}\gg t_{accr}$, one concludes that the radiation is traped inside the spiral flow on the accretion time scale and, therefore, this flow is radiatively inefficient. From the point of view of an observer, which detects the softer part of the spectrum of outgoing radiation (below $\sim$ eVs), the MAD will look like a Shakura-Sunyaev disk truncated at the inner radius $R_{tr}$, which coinsides with the transition radius between the inner magnetically dominated region and outer axisymmetric accretion disk. Typically, in observations, $R_{tr}$ is in the interval from a few tens to hundreds of $R_g$ (e.g., in Cyg X-1, see Done \& Zycki 1999), which is consistent with that obtained in our Models~A and B. The observed specta of black hole binaries in the low/hard state are dominated by the hard x-ray component (e.g., Done, Gierli\'nski, \& Kubota 2007) and this can be explained by the radiation from the hot, optically thin magnetized medium, which surrounds the accreting spiral flows and in which the binding energy of these flows is released. The synchrotron radiation from the magnetized medium and Poynting jets in the MAD (see Goldston, Quataert, \& Igumenshchev 2005) can be used to explain the observed radio luminosity in the low/hard state; this luminosity is believed to be due to steady jets (e.g., Corbel et al. 2003; Gallo, Fender, \& Pooley 2003). Our simulations assume accretion disks around black holes and can be relevant to objects with relativistic jets containing accreting stellar-mass black holes (e.g., in micro-quasars; black holes resulted from type Ib/c supernova explosions and mergers of two compact objects) and supermassive black holes (in galactic centers). However, we believe that our main results can also be relevant to nonrelativistic astrophysical objects, in which accretion disks and jets are observed. These objects include, for example, young stellar objects and accreting stars (e.g., white dwarfs) in binary systems. Qualitatively, we expect that the structure and dynamics of MAD's and Poynting jets are similar in the both relativistic and nonrelativistic cases. We expect, however, large quantitative differences in these two cases because of the different energy-density scales involved in the regions of jet formation. Black holes, which are capable of launching jets almost from the event horizon, can produce ultra relativistic jets (e.g., McKinney 2006). Jets from nonrelativistic objects, which have the surface radius $R_*\gg R_g$, are limited by the velocities $v\sim\sqrt{GM/R_*}\ll c$. For example, the latter formula gives the upper estimate of the jet velocity $\sim 400\,km/s$ from the solar-type stars. The problem of inward transport and amplification of the vertical field in turbulent accretion disks was intensively discussed in past and recent years (see, e.g., Spruit \& Uzdensky 2005). The solution of this problem can help to discriminate models of accretion disks, which are consistent with observations (e.g., Meier \& Nakamura 2006; Schild, Leiter, \& Robertson 2006). Our simulation results show that the vertical field is transported inward and amplified independent of the disk structure, either laminar or turbulent. It is worth noting, however, that magnetic fields in our models are imposed and relatively large. The assumed strength of these fields exceeds the possible strength of the self-sustained magnetic fields that could be developed due to the MRI (Sano et al. 2004). Therefore, these results should be considered with some caution, because they do not represent the case of weak vertical magnetic fields. The strong vertical field in the center of accretion disks can be, in principal, a relic field that is inherited from the previous evolution. This field can appear, for example, in the merger scenario (merger of two magnetized neutron stars or a black hole with a magnetized neutron star, e.g., Berger et al. 2005) or in the course of the gravitational collapse of an extended ``proto" object (e.g., proto-stellar cloud, supernova progenitor), which produces a significantly more compact object (protostar, black hole). In the latter case, the proto object may contain some amount of the poloidal field, which will be amplified and accumulated at the center during the collapse. After the formation of the compact object, the remained noncollapsed mass can still move inward, forming an accretion disk and confining the field in the vicinity of the object. Depending on the relative strength of this relic field and the mass accretion rate, MAD's and Poynting jets can be developed. The considered scenario can be applied to young stellar objects (T-Tauri stars, e.g., Donati et al. 2007) and the hyper-accretion model for gamma-ray bursts (Woosley 1993; Paczy\'nski 1998). The spiral-flow pattern in MAD's rotates with about the same angular velocity at all radial distances; i.e., it rotates almost as a rigid body. Such a rotation can result in quasi-periodic oscillations (QPO's) in the emitted radiation, if the disk's axis is inclined to the line of view of an observer (e.g., Alpar \& Shaham 1985; Lamb et al. 1985; Strohmayer et al. 1996; Lamb \& Miller 2001; Titarchuk 2003). The frequency of these QPO's should be related to the rotation of the spiral pattern, which angular velocity is defined by the radius of the magnetically dominated region and can be a fraction ($\sim 0.5-1$) of the orbital velocity at this radius. More investigations are required to make quantitative predictions about QPO's from MAD's. \acknowledgments This work was supported by the U.S. Department of Energy (DOE) Office of Inertial Confinement Fusion under Cooperative Agreement No. DE-FC52-92SF19460, the University of Rochester, the New York State Energy Research and Development Authority. \clearpage
{ "timestamp": "2007-11-28T01:01:40", "yymm": "0711", "arxiv_id": "0711.4391", "language": "en", "url": "https://arxiv.org/abs/0711.4391" }
\section{Introduction} The discovery of irregularities in the cosmic ray energy spectrum at the energy of $\sim 3 \times 10^{15}$~eV (Khristiansen et al., 1956~\cite{1}) and $\sim 8 \times 10^{18}$~eV (Krasilnikov et al., 1978~\cite{2,3,4}), the detection of sharp decreases in the cosmic ray intensity at $E_{0} > 5 \times 10^{19}$~eV (Greisen-Zatsepin-Kuzmin effect, 1966~\cite{5, 6}) at the EAS arrays in Yakutsk, HiRes (USA), AUGER (Argentina) are the most important achievement in the investigation of the superhigh and ultrahigh energy cosmic rays in recent years. Such a character of spectrum turn out to be associated directly with processes in interstellar space, namely, with the origin, acceleration and propagation of cosmic rays in the Galaxy and beyond. The interpretation of these experimental facts using the different models of cosmic ray origin still remains to be answered. In this paper the comparison of the cosmic ray energy spectrum by EAS Cherenkov light measurements at the Yakutsk array~\cite{7, 8} with the calculations according to an anomalous diffusion model of cosmic rays in interstellar space~\cite{9} is performed. \section{Method to construct the EAS spectrum} The showers at the Yakutsk array are selected with the central register by both scintillation and Cherenkov ``masters''~\cite{10, 11}. The all showers registered form the database of the Yakutsk EAS array. To construct the spectrum in energy, scattered by EAS particles in the atmosphere (Cherenkov radiation) the following selection criteria of showers are used: a) a shower core is to be placed within a perimeter of the array for the giant showers and near a center of the array for the showers with $E_{0} < 10^{18}$~eV. The showers whose cores are near the observation station $R \le 60$~m are excluded from a sampling: b) the probability to register a shower by Cherenkov photons is $W_{\text{ch}} \ge 0.9$; c) a zenith angle is less than one-half of an aperture of Cherenkov detector, i.e. $\theta < 55^{\text{o}}$ in the case of the detector of the first type and $\theta < 60^{\text{o}}$ for the second type detector; d) a transmission coefficient of the atmosphere is $\ge 0.60$ for the wave length of $430$~nm. Thus, more than 60000 showers with $E_{0} \ge 10^{17}$~eV and 300000 showers with $10^{15} \le E_{0} \le 10^{17}$~eV were recorded in the database. To construct the spectrum, the showers were selected by the classification parameter $Q(R=150)$, i.e. by Cherenkov light flux density at a distance 150~m from a core, which was proportional to the primary shower energy. The measurement accuracy for $Q(R=150)$ in the individual showers was $\delta = \Delta Q(R=150) / Q(R=150) \ge 15$\%. The estimation of the shower energy $E_{0}$ is determined by a quasicalorimetric method which does not depend on the EAS development model. A basis of the method is experimental data about the Cherenkov light total flux, $F$, the total number of charged particles, $N_{\text{s}}$, and the total number of muons with $E_{\text{thr}} \ge 1~GeV$, $N_{\mu}$~\cite{12, 13, 14}. The energy of individual showers is determined by the following formula: \begin{equation} E_{0} = (9.1 \pm 2.2) \cdot 10^{16} \times Q(R=150)^{0.99 \pm0.02} \label{eq1} \end{equation} The intensity of cosmic ray flux in the given interval of EAS classification parameter is found as a ratio of the number of registered EAS events to $S_{\text{eff}} \cdot T \cdot \Omega$. \section{Results and Discussion} The differential energy spectrum of primary cosmic rays in the interval of $10^{15} - 5 \times 10^{19}$~eV obtained from a totality of the all Cherenkov detector measurement data at the Yakutsk EAS array is shown on the Fig.~\ref{fig01}. Our data confirm an irregularity of the spectrum of ``knee'' type in the energy range of $(2-5) \times 10^{15}$~eV discovered in~\cite{1}, and the irregularity of ``ankle'' type at $E_{0} \sim 8 \times 10^{18}$~eV. It is established that in the first case the spectrum index is $\gamma = 2.7 \pm 0.1$ below the break and $\gamma = 3.03 \pm 0.05 at E_{0} > 3 \times 10^{15}$~eV, and in the second case, the more sloping spectrum with $\gamma = 2.6 \pm 0.3$ at $E_{0} > 8 \times 10^{18}$~eV is observed. \begin{figure}[ht] \centering \includegraphics[width=0.42\textwidth,clip]{fig01} \caption{Energy spectrum of primary cosmic rays by measurement data of Cherenkov light at the Yakutsk complex EAS array.} \label{fig01} \end{figure} \begin{figure} \centering \includegraphics[width=0.42\textwidth,clip]{fig02} \caption{Differential cosmic ray intensity versus the energy. The points are Yakutsk array data, curves are the calculation from~\cite{9}.} \label{fig02} \end{figure} For the period of continuous observations of Cherenkov radiation more than 30 years (10\% relative to one year time of EAS registration with the scintillation detectors), the showers with $E_{0} > 6 \times 10^{19}$~eV did not detect. This fact confirms once more the GZK hypothesis~\cite{5, 6} about the sharp break in the cosmic ray energy spectrum at $E_{0} > 5 \times 10^{19}$~eV. \subsection*{The galactic model} The attempt to explain a form of obtained spectrum from the point of view of cosmic ray anomalous diffusion model and fractality of the Galaxy's magnetic field was made by Lagutin et al~\cite{9}. The basis for the cosmic ray propagation in the Galaxy is the following assumptions: a) after the generation in the sources, the particles move in fractal interstellar medium by means of two ways: the first way is ``Levy flights'', the second way is the motion along a spiral in the nonhomogeneous magnetic field, $b$) the particles exist during anomalous long time. The lifetime of particles is of a wide distribution and its tail is described by a power law $q(t) \propto B \cdot t^{-\beta -1}, t \to \infty, \beta < 1$ (so-called ``Levy trapping time''). Calculations of the spectrum were separately made for each of following groups of nuclei: p, He, CNO, N-Si, Fe. The resulting sum spectrum for the all particles is shown by a solid curve in Fig.~\ref{fig01}. From the calculations it follows that the suggested model reproduces the irregularity in the energy spectrum of the ``knee'' type at $E_{0} \simeq 3 \times 10^{15}$~eV and also the irregularity of the ``ankle'' type at $E_{0} \simeq 8 \times 10^{18}$~eV. This model does not explain the behavior of a spectrum in the energy region of $10^{17} - 10^{18}$~eV and the break of the spectrum at $E_{0} > 6 \times 10^{19}$~eV in more detail. The mass composition in the energy region of $5 \times 10^{15} - 5 \times 10^{18}$~eV is some heavier than at $E_{0} \simeq 10^{19}$~eV, but this change is not very significant that is expected from an experiment (see Fig.~\ref{fig03}). \subsection*{The galactic model with the sources of two types} In contrast to~\cite{9}, in the paper~\cite{15} a scenario is considered, in which supernovae are the main sources of cosmic rays and the acceleration up to $E_{\text{max}} \simeq 105 \cdot Z$~GeV takes place in the shock fronts. The particle spectrum formed in this case can be presented in the form of $S_{\text{SN}} \sim E^{-2} \theta (E_{\text{max}} - E)$, where the Heaviside function $\theta(x)$ reflects qualitatively the presence of a sharp cut-off in the spectrum at $E > E_{\text{max}}$~\cite{16, 17}. The new calculations fulfilled by the above scenario of particle generation in the sources of two different types under the assumption of anomalous diffusion of particles in inhomogeneous medium show that at some parameters the anomalous diffusion model describes satisfactorily the features of cosmic ray energy spectrum and mass composition up to $E_{0} \sim 10^{18}$~eV observed in an experiment. First of all, it refers to the fine structure of cosmic ray intensity change depending on the energy (see Fig.~\ref{fig01}). By using these calculations, the sharp peaks in the mass composition depending on energy are also observed (see Fig.~\ref{fig03}). In this connection, it is of interest to compare calculations in mass composition with experimental data obtained at the Yakutsk EAS big and small Cherenkov sub-arrays in recent years. \subsection*{Mass composition of primary cosmic rays} Fig.~\ref{fig03} presents the results in mass composition of primary cosmic rays of the Yakutsk array. The data were obtained in the framework of QGSJET-01 model and two-component mass composition (proton-iron nucleus). The several characteristics corresponding to the radial and longitudinal development of EAS are used in the analysis~\cite{18, 19, 20, 21, 22}. \begin{figure} \centering \includegraphics[width=0.42\textwidth, clip]{fig03} \caption{Mass composition of cosmic rays at superhigh and ultrahigh energies. The curve is a calculation by Lagutin et al (2001) according to the anomalous diffusion model for cosmic ray propogation.} \label{fig03} \end{figure} \begin{figure} \centering \includegraphics[width=0.42\textwidth, clip]{fig04} \caption{Comparison of the experimental spectrum with the calculated spectrum from~\cite{23} for the metagalactic protons ($E_{0} > 5 \times 10^{17}$~eV) and galactic iron nuclei ($E_{0} = 10^{15} - 5 \times 10^{17}$~eV).} \label{fig04} \end{figure} The value $\left<\ln{A}\right>$ in each case is determined by using the interpolation method~\cite{24}. It is seen from Fig.~\ref{fig03} that the mass composition is varied up to heavy elements in the energy region of $(2-5) \times 10^{17}$~eV and becomes more light beginning with $E_{0} \sim 3 \times 10^{18}$~eV. The lines are the calculations according to the anomalous diffusion model for the propagation of cosmic rays in the Galaxy (Fig.~\ref{fig03}, two sources) in the case of inhomogeneous galactic medium. In the first case, the monotone change in the mass composition up to $E_{0} \ge 3 \times 10^{18}$~eV is observed, after of which the mass composition becomes more light. In the second case, the complicated structure in the dependence of mass composition on the energy is observed, peaks for the nuclei of different mass in the energy region of $10^{15} - 10^{17}$~eV are noticeable. According to a hypothesis~\cite{25} and calculations from \cite{15}, such an inhomogeneous structure can be formed by a near supernova. Our data (Fig.~\ref{fig03}) do not contradict to this hypothesis. Such a sharp change of the mass composition in the energy region of $5 \times 10^{16} - 5 \times 10^{17}$~eV is not explained in the framework of the galactic model and is likely associated with the existence of a transition boundary from galactic to metagalactic cosmic rays. This conclusion is confirmed by calculations from~\cite{23}, where a scenario of galactic and metagalactic origin of cosmic rays is considered. These calculations are shown in Fig.4 together with our experimental data. It can be seen from Fig.~\ref{fig04} that cosmic rays in the energy region of $5 \times 10^{16} - 5 \times 10^{17}$~eV are most likely of galactic origin with the noticeable portion of heavy nuclei in the total flux. It should be noted the estimations of cosmic ray mass composition in the region after the ``knee'', obtained at the compact arrays, agree well with each other. The same cannot be said of the energy region of $\sim 10^{18}$~eV (see Fig.~\ref{fig03}) where HiRes array data point to more speedy enrichment of primary radiation by the light nuclei and protons as compared with the Yakutsk array data. The Yakutsk EAS array data, on the contrary, show the gradual change from the heavy to light composition (protons and He nuclei) at $E_{0} \sim 10^{19}$~eV. In both cases, data point to the existence of the tendency of ``protonization'' of primary cosmic rays at $E_{0} >3 \times 10^{18}$~eV. \subsection*{Conclusions} Direct measurements of the cosmic ray energy spectrum in the region of ultrahigh energies (in energy scattered by EAS particles in the atmosphere) have confirmed the complicate form of spectrum. The spectrum becomes steeper at $E_{0} \ge 3 \times 10^{15}$~eV and more sloping at $E0 \ge 8 \times 10^{18}$~eV. A character of energy dependence of $\left<\ln{A}\right>$ by the Yakutsk EAS data point to the change of the mass composition of primary particles at singular points of cosmic ray energy spectrum. The value $\left<\ln{A}\right>$ rises with the energy after the “knee” up to its maximum equal to $3.5$ at $(2-5) \times 10^{17}$~eV and then it begins to decrease. Such an energy dependence of $\left<\ln{A}\right>$ does not contradict a hypothesis of cosmic rays propagation according to laws of the anomalous diffusion model in fractal interstellar medium (Lagutin et al., 2001). The value $\left<\ln{A}\right>$ at $E_{0} > 10^{18}$~eV decreases gradually and at $E_{0} \sim 10^{19}$~eV the mass composition consists of He nuclei and protons. The cosmic ray intensity beyond $E_{\text{thr.}} > 6 \times 10^{19}$~eV decreases sharply and this effect is not described in the framework of the galactic model only. Such a character of spectrum does not contradict to the calculations by Berezinsky et al~\cite{23} for the metagalactic model, in which the ``ankle'', observed in the experiments on ultrahigh energy cosmic ray registration, can be produced by the proton component only arriving from the Metagalaxy. Thereby, the details of experimental spectrum form, for example, ``dip'', i.e. the decrease of intensity at $E_{0} \times 10^{19}$~eV, are caused by, most likely, the interaction of extragalactic protons with a relic radiation photons ($p + \gamma_{\text{\footnotesize{}CMB}} \to \text{p} + e^{+} + e^{-}$). As a direct argument of this hypothesis, the anisotropy can be used which is related to the origin and sources of cosmic rays. Based on data of~\cite{26, 27, 28}, at $E_{0} \ge 8 \times 10^{18}$~eV the weak correlation in the arrival directions of EAS with the Galaxy plane and the close correlation with the Supergalaxy plane are observed and that the quasars can be the possible sources of cosmic rays.
{ "timestamp": "2007-11-16T03:50:04", "yymm": "0711", "arxiv_id": "0711.2548", "language": "en", "url": "https://arxiv.org/abs/0711.2548" }
\section{Introduction} Field-induced magnetic ordering (FIMO) in spin-gapped systems, in which an energy gap exists for low-lying excited states, has been investigated in a vast number of compounds, particularly in the context of the Bose-Einstein condensation (BEC) of triplet magnons \cite{nikuni}. The BEC picture is useful for understanding the nature of the FIMO with the commensurate (C) antiferromagnetic order pependicular to the field direction. Recently, Suzuki {\it et al.} and Maeshima {\it et al.} have added a new aspect to the FIMO on the basis of numerical analyses combined with field theories \cite{suzuki, maeshima}. These authors have predicted that a magnetic field induces a novel incommensurate (IC) order parallel to the field direction in $S=1/2$ alternating chains with a frustrated next-nearest-neighbor (NNN) interaction. Around the central field of the field-induced Tomonaga-Luttinger liquid (TLL) phase of this system between the lower and upper critical field, $H_{\rm C1}$ and $H_{\rm C2}$, frustration changes the dominant spin correlation from C to IC. If small inter-chain interactions exist, the dominant IC correlation leads to long-range IC ordering in the field direction. In the case frustration is not strong enough to stabilize the IC order at high temperatures, a first order phase transition will happen from the BEC to the IC order at very low temperatures\cite{maeshima, maeshima2}. The theoretical studies mentioned above \cite{suzuki, maeshima} have been stimulated by experimental works on the organic radical compound pentafluorophenyl nitroxide (F$_5$PNN) \cite{across, izumi, izumithesis, lt23}. The magnetism of F$_5$PNN arises from unpaired electrons delocalized around the NO moieties. Although this compound has a uniform chain structure at room temperature, the magnetic susceptibility and the magnetization curve at low temperatures are well reproduced by calculations for an $S=1/2$ alternating chain model which is described by the spin Hamiltonian \cite{across}; \begin{equation} H = -2J \sum_{i}^{N/2} (S_{2i-1}\cdot S_{2i}+\alpha S_{2i}\cdot S_{2i+1}). \end{equation} Here, $S$ denotes the $S=1/2$ Heisenberg-type spin operator, $N$ is the total number of spins, and $\alpha$ is the alternation ratio between competing two nearest-neighbor interactions in a one-dimensional chain. When $\alpha$=1, the system becomes a uniform chain, whereas when $\alpha $=0 the system breaks up into the assembly of isolated dimers. In Ref. \onlinecite{across}, the alternation ratio $\alpha=0.4$ and exchange interaction $2J/k_\mathrm{B}=-5.6$ K were obtained for F$_5$PNN. The lower and upper critical fields of F$_5$PNN are determined to be about $H_\mathrm{C1}=3.0$ T and $H_\mathrm{C2}=6.5$ T from the magnetization curve. NMR shows a TLL behavior in spin-lattice relaxation and provides evidence for a NNN interaction \cite{izumi, suga, izumithesis}. In previous works, we observed FIMO by measuring the specific heat of a polycrystalline sample in magnetic fields up to 8.0 T ( $>H_\mathrm{C2}$ ) \cite{prl}. Above the critical temperature of the FIMO, the temperature dependence of the specific heat $C(T)$ in magnetic fields was in good qualitative agreement with a numerical calculation which assumes the TLL \cite{prl, wang}. In this paper, we present the $H$-$T$ phase diagrams of a single crystal and powder of F$_5$PNN obtained from detailed specific heat measurements in magnetic fields. Reentrant $H$-$T$ phase diagrams for the FIMO phase are obtained for both samples. However, the shape of the phase boundary depends on the form of the sample. That of the single crystal is symmetric with respect to a central field of the gapless field region between $H_\mathrm{C1}$ and $H_\mathrm{C2}$, whereas the powder has a phase boundary which is distorted and pushed to lower temperatures than that of the single crystal. \section{Experimental procedures} F$_5$PNN was prepared using the method described in Ref. \onlinecite{sample}. Specific heat measurements were performed by the adiabatic heat-pulse method using a $^3$He-$^4$He dilution refrigerator. The powder sample was mixed with Apiezon N grease to ensure good thermal contact, and was mounted on the sample cell in the refrigerator. The single crystal sample was attached to the cell with the same grease. The nuclear contributions of hydrogen and fluorine to the specific heat were subtracted. \begin{figure}[t] \begin{center} \includegraphics[width=8cm]{single031107.eps \end{center} \caption{(color online) Specific heat of the single crystal in magnetic fields. Arrows indicate peak temperatures of the FIMO. Upper panel: 5.0 T $\leq H\leq$ 6.5 T. Lower panel: 2.5 T $\leq H\leq$ 4.5 T. } \end{figure} \begin{figure}[t] \vspace{0.18cm} \begin{center} \includegraphics[width=8.3cm]{BpowderBUT5p2T031107.eps \end{center} \vspace{0.25cm} \caption{(color online) Specific heat of the powder in magnetic fields. Arrows indicate peak temperatures of the FIMO. Upper panel: 4.75 T $\leq H\leq$ 6.75 T. Lower panel: 2.5 T $\leq H\leq$ 4.5 T. } \end{figure} \section{Results} Figure 1 shows $C(T)$ of the single crystal in magnetic fields. A sharp peak due to the FIMO is clearly seen in fields between 3.25 T and 6.0 T. A small peak is observed at 3.0 T, and an upturn indicating a peak at a lower temperature at 6.25 T. Also, exponential temperature dependences are observed at 2.5 T and 6.5 T, indicative of energy gaps for low-lying excitations. We conclude from these observations that the critical fields of the single crystal are $H_\mathrm{C1}\simeq 3.0$ T and $H_\mathrm{C2}\simeq 6.25$ T. A sharp peak in $C(T)$ is clearly observed also in the powder in fields $3.25$ T $\leq H\leq$ 6.0 T as shown in Fig. 2, and the peak temperatures are in accordance with those of our previous results \cite{lt23, prl}. However, the field and temperature dependences of the peak are quite different from those in the single crystal. As the field increases from 3.5 T, the peak becomes much sharper. The field dependence of the peak temperature is weaker than for the single crystal. At 3.0 T, 6.25 T, and 6.5 T, a sharp upturn is observed indicating a peak at lower temperatures, and exponential behaviors are observed at 2.5 T and 6.75 T. Based on these features, the gapless field region of the powder is most likely 3.0 T $\leq H\leq$ 6.5 T. This field region is wider than that of the single crystal. Figure 3 is the $H$-$T$ phase diagram of the single crystal and powder obtained from the peaks in the specific heat. We note two differences between the $H$-$T$ phase boundaries for the two sample forms. One is that the peak temperatures are lower for the powder than for the single crystal. The other is a difference in the shape of the phase boundary between the FIMO and paramagnetic phase. The phase boundary of the single crystal is symmetric with respect to the central field of the gapless field region as observed or expected in isotropic spin-gapped compounds investigated so far \cite{FIMOs}. In contrast, that of the powder is distorted. Since the peak in $C(T)$ of the powder is sharp even at 6.0 T in Fig. 1(a), it is unlikely that the distinct phase boundary of the powder originates from anisotropy effects. In addition, it is revealed by high-field ESR measurements on the powder sample of this compound that the g-value is almost 2.0 \cite{ESR}. \begin{figure}[tbp] \includegraphics[width=8cm]{pdsmall031107.eps \caption{(color online) Magnetic field versus temperature phase diagram of the single crystal and powder of F$_5$PNN obtained from the specific heat measurements. Open and filled circles are peak positions of the specific heats of the single crystal and powder, respectively. Solid and broken lines are guides for the eye.} \end{figure} \section{DISCUSSION} The observed distorted phase boundary of the FIMO is similar to that of $S=1/2$ strongly frustrated alternating chain models \cite{maeshima}. The models exhibit a first-order phase transition at very low temperatures from a conventional field-induced antiferromagnetic order of the spin components perpendicular to the external field direction, which is interpreted as the BEC of triplet magnons, to an IC order along the field direction around the middle of the gapless field region where the IC correlation is dominant. Because frustration suppresses transverse fluctuations, and then decreases the antiferromagnetic ordering temperature in this field region, the phase boundary for the FIMO is distorted. To argue the possibility that an IC order is realized in the powder, we must first examine if frustration is necessary to explain the powder result. \begin{figure}[t] \begin{center} \includegraphics[width=8.5cm]{singleanddmrg2J.eps \end{center} \begin{center} \includegraphics[width=8.5cm]{powderanddmrg2J.eps \end{center} \caption{Temperature dependence of the magnetic specific heats $C_\mathrm{m}(T)$ of F$_5$PNN single crystal and powder at zero field, together with calculated specific heats with three sets of values for the exchange interaction $J/k_\mathrm{B}$ and alternation ratio $\alpha$ based on the finite temperature DMRG. Upper panel: the $C_\mathrm{m}(T)$ of the single crystal and calculation with $2J/k_\mathrm{B}=-5.6$ K and $\alpha=0.4$. Lower panel: the $C_\mathrm{m}(T)$ of the powder and calculations with $2J/k_\mathrm{B}=-5.6$ K and $\alpha=0.4$, $2J/k_\mathrm{B}=-5.6$ K and $\alpha=0.6$, and $2J/k_\mathrm{B}=-6.8$ K, $J'/J=0.2$ and $\alpha=0.7$. } \end{figure} To determine the exchange interactions $J$ and alternation ratios $\alpha$ of the single crystal and powder, we examine the magnetic specific heat at zero field for both samples. The lattice contribution to the total specific heat is estimated from the data at zero field so that the total magnetic entropy for $N$ spins will approach $Nk_\mathrm{B}\mathrm{ln}(2S+1)$ at high temperatures where the magnetic susceptibility $\chi $ times temperature $T$ approaches the value for an $S=1/2$ system. The results are compared with numerical calculations based on the finite temperature density matrix renormalization group (DMRG) \cite{DMRG} as shown in Fig. 4. The upper panel of Fig. 4 shows the single crystal result and a calculation with the set of parameters $2J/k_\mathrm{B}=-5.6$ K and $\alpha=0.4$, which have been obtained from the magnetic susceptibility and magnetization of a single crystal \cite{across}. The quantitative agreement between the experimental and numerical results means that frustration in the single crystal is too small to detect in the specific heat if it exists. In the lower panel of Fig. 4, we compare the result of the power with numerical calculations with various parameter sets. It should be noted that the calculation with $2J/k_{\rm B}=-5.6$ K and $\alpha=0.4$, which well reproduces the single crystal result, is largely different from the powder result implying the parameters of the powder are not equal to those of the single crystal. Although results for $2J/k_{\rm B}=-5.6$ K and $\alpha=0.6$ are better than those for the first parameter set, clear differences appear in the both side of the peak temperature ($\sim 2$ K). Finally, our best result is obtained by assuming a NNN interaction for $J/k_{\rm B}=-6.8$ K, $\alpha=0.7$ and $J'/J=0.2$. We note that the enhanced $J$ and $\alpha$ explain the wider gapless field region of the powder because $J$ and $\alpha$ govern the width of the gapless field region of $S=1/2$ bond-alternating chains \cite{bonner}. Also, this agreement rules out the possibility that the distinct phase boundary of the powder is ascribed to the disappearance of magnetic moments which comes from the sample deterioration. \begin{figure}[tbp] \begin{center} \includegraphics[width=8cm]{incommephase2.eps \end{center} \caption{Alternation ratio $\alpha$ versus NNN interaction $J'/J$ phase diagram at zero temperature at the half value of the saturation magnetization, which is equivalent to the 1/2 plateau phase diagram in Ref. \onlinecite{tonegawa}. The filled circle denotes the set of parameters obtained for the F$_5$PNN powder in this study.} \end{figure} The next thing to do is to check whether the set of parameters for the powder is comparable to those in which an IC order is theoretically predicted to appear. Figure 5 shows the different regions of the dominant correlation at the half value of the saturation magnetization in the frustrated alternating chain model as a function of $J'/J$ and the alternation ratio $\alpha$ at $T=0$. The IC correlation becomes dominant in the same region where the half-magnetization plateau is stable \cite{maeshima2}. The set of parameters for the F$_5$PNN powder, $J'/J=0.2$ and $\alpha=0.7$, turns out to be in this region, shown as a filled circle in this figure. This result strongly suggests that the IC correlation is dominant in the powder around the center field of the gapless field region and an IC order exists at very low temperatures. However, there remains a question why the NNN interaction exists only in the powder. The large pressure dependences of the magnetic susceptibility and specific heat of F$_5$PNN reported in previous works give us a possible answer to this question \cite{hoso2, mito}. According to these works, $\alpha $ and $J$ increase with increasing external pressure, and even at $P=0$, mixing powder F$_5$PNN with Apiezon N grease changes these values. The grease solidifies at low temperatures and gives some stress to the powder inside the solid. Effective pressure by the solidification of the grease is also reported for the powder of another organic compound \cite{mukai}. Generally, an external pressure enhances inter-chain interactions which increase the ordering temperature of the FIMO. Nevertheless, the ordering temperatures of F$_5$PNN is higher for the single crystal than for the powder which can be under pressure as mentioned above. The strength of an antiferromagnetic interaction in organic magnetic materials depends on how the molecular orbital of an unpaired electron overlaps with the others. Since this orbital spreads rather widely in each molecule, the small variation in the molecular stacking can change the magnetic property drastically \cite{f2pnnno}. From this point of view, an external pressure most likely changes the molecular stacking in F$_5$PNN so that the frustrated NNN interaction, which suppresses the ordering temperature, will be enhanced much more than the inter-chain interactions. Very recently, we have seen a more clearly distorted phase boundary for the FIMO around the central field in the specific heat measurement of deuterated F$_5$PNN powder sample. This result will appear somewhere else. To investigate quantitatively the pressure-induced frustration in this compound, we have proceeded specific heat measurement in magnetic fields under pressure. \section{Summary} We have performed detailed specific heat measurements on the $S=1/2$ alternating chain material F$_5$PNN in magnetic fields using a single crystal and powder. The shape of the phase boundary for the field-induced magnetic ordered phases is different between the two sample forms. We have shown the possibility of the pressure-induced frustration in the powder which should lead to field-induced incommensurate ordering around the central field besides the Bose-Einstein condensation of triplet magnons, by quantitatively comparing zero-field magnetic specific heats of two samples with numerical calculations based on the finite temperature density matrix renormalization group. A future challenge is the direct observation of the incommensurate ordering. \section{Acknowledgements} We thank Yasumasa Takano for helpful advice and valuable discussions. We are grateful to Kazuyoshi Takeda, Masaki Mito, Seiichiro Suga, Takafumi Suzuki, Akinori Tanaka, Toshihiro Idogaki, Kiyohide Nomura, Luis Balicas, and Takahiro Sakurai for helpful comments. Y.Y. was supported by Japan Society for the Promotion of Science.
{ "timestamp": "2007-11-15T06:12:07", "yymm": "0711", "arxiv_id": "0711.2336", "language": "en", "url": "https://arxiv.org/abs/0711.2336" }
\section{Some features of Ultraperipheral Collisions (UPC)} Photon-photon and photon-hadron interactions can also be studied in hadron-hadron collisions~\cite{url}. This may be surprising since in general such collisions are dominated by strong interactions between the hadrons. However, by choosing collisions with large impact parameter b (or, equivalently, small momentum transfer) one can suppress these strong interactions. The time-dependent electromagnetic field of a fast moving charged particle can be thought of as a spectrum of (quasireal, or equivalent) photons~\cite{fermi}, see Figure \ref{Fig:PH}. The determination of the equivalent (or Weizs\"acker-Williams) photon spectrum corresponding to a fast particle moving past an observer on a straight line path with impact parameter $b$ is a textbook example ~\cite{jac}. The probability $P(b)$ of a specific photon-hadron reaction to occur in a collision with an impact parameter $b$ is given by $P(b)=N(\omega, b) \sigma_{\gamma h}(\omega)$, where $\sigma_{\gamma h}$ is the corresponding photoproduction cross section. The equivalent photon spectrum can be calculated analytically, a useful approximation for qualitative considerations is \begin{equation} N(\omega, b)=\frac{Z^2 \alpha}{\pi^2 b^2} \label{Eq:nb} \end{equation} for $\omega<\frac{\gamma}{b}$ and zero otherwise. The nuclear charge is given by $Z$, heavy ions have particularly high photon fluxes, however, this is partially offset by the lower ion-ion luminosities, as compared to the p-p case. \begin{figure}[h] \centerline{\includegraphics[width=1.0\columnwidth]{f1.eps}} \caption{A fast charged particle moving on a straight line with impact parameter b causes a time-dependent electromagnetic field at the point of the observer. This field corresponds to a spectrum of equivalent photons.}\label{Fig:PH} \end{figure} The impact parameter b is restricted to \begin{equation} b>b_{min} \sim R_1 + R_2 \end{equation} where $R_1$ and $R_2$ denote the sizes of the hadrons. For heavy ion scattering the Coulomb parameter $\eta \equiv \frac{Z^2 e^2}{\hbar v}\sim Z^2/137$ is much larger than unity and it is in principle possible to determine the impact parameter by measuring the angle of Coulomb scattering. Whereas this is experimentally feasible at lower ($\sim GeV/A$) energies ~\cite{aum}, this angle is too small at collider energies. So one generally measures quantities integrated over all impact parameters. Too small impact parameters are recognized since the event is dominated by the violent strong interactions. The photon spectrum Eq.~\ref{Eq:nb} extends up to a maximum photon energy given by \begin{equation} \omega_{max}=\frac{\gamma}{b_{min}} . \end{equation} This energy is about 3 GeV at RHIC (Au-Au, $\gamma \sim 100$), and 100 GeV at LHC (Pb-Pb, $\gamma \sim 3000$) in the collider system. \section{Multiphoton processes: a possible trigger on UPC} For heavy ions the probability of an electromagnetic interaction in ultraperipheral collisions is especially large, and multiphoton processes occur, see e.g.~\cite{npa}. We mention $e^+e^-$ pair production where the impact parameter dependent total pair production probability $P(b)$ is of order unity. Multiple pairs can be produced, however they may be hard to detect due to their low transverse momentum. The nuclear giant dipole resonance is excited with probabilities of order of one third. In Figure \ref{Fig:mua} one of the graphs is shown which leads to the electromagnetic production of a $\rho^0$ along with the excitation of the giant dipole resonance. These graphs can conveniently be evaluated in semiclassical or eikonal theories~\cite{npa}. \begin{figure}[h] \centerline{\includegraphics[width=1.0\columnwidth] {baur_gerhard.fig2.ps}} \caption{A graph contributing to the simultaneous production of a $\rho$-meson and the excitation of the giant dipole resonance (GDR).}\label{Fig:mua} \end{figure} The giant dipole resonance decays dominantly into a neutron. This neutron is detected in the forward direction and can serve as a trigger on UPC. \section{UPC at RHIC} The physics of UPC at RHIC and results from the STAR detector were covered by J. Seger in the session on photon- and electroweak boson physics, from HERA, RHIC and Tevatron to LHC. A unique feature to photoproduction in hadron-hadron collisions is an interference effect ~\cite{kn}: a vector meson can be produced by a photon originating from either of the hadrons. It was shown in~\cite{kn} that this interference effect leads to a reduction of the transverse momentum spectrum of the vector mesons for small transverse momenta. Another theoretical approach~\cite{hbt} leads to very similar conclusions. (Preliminary) experimental results from STAR/RHIC indeed show a dip for small transverse momenta, see e.g. Ref. \cite{yr}. \section{Opportunities for UPC at LHC} The maximum photon energy scales linearly with the Lorentz factor $\gamma$, see eq. 3. This leads to a significant widening of the opportunities at LHC as compared to RHIC. A most promising area is low-x QCD studies. The experiments at HERA have shown that photoproduction processes provide a well-understood probe of the gluon density in the proton. At LHC, such processes could be extended to invariant $\gamma p$ energies exceeding the maximal HERA energy by a factor of 10. This would allow to use dijet (charm, etc.) production to measure the gluon density in the proton and/or nucleus down to $x \sim 3 \times 10^{-5}$. Ultraperipheral collisions would also allow one to study the coherent production of heavy quarkonia, $\gamma + A \rightarrow J/\Psi (\Upsilon) + A$ at $x \lessapprox 10^{-2}$, and to investigate the propagation of small dipoles through the nuclear medium at high energies, see Ref. \cite{fra}, see also Refs.~\cite{kopel,macha}. Dijet production via photon-gluon fusion is calculated in Ref. \cite{svw}. Very large rates are obtained that will considerably extend the HERA x range. In this session plans for studying UPC physics with heavy ions at the LHC were covered by J.Nystrand (ALICE), D.D'Enterria (CMS), and V. Pozdnyakov (ATLAS). In addition to diffractive processes in proton-proton collisions at LHC also a rich program of proton-photon and photon-photon physics can be pursued, see Ref. \cite{cern}. The photon flux is lower as compared to the heavy ion case due to the $Z^2$-factor, but this is at least partly compensated by higher beam luminosities. The photon spectrum is harder due to the smaller size as compared to the heavy ions, this leads to a lower value of $b_{min}$ in Eq. 3 . Possibilities for electroweak physics and beyond were presented by S.Ovyn ($\gamma p$) and T. Pierzchala ($\gamma \gamma$) in this session. Tagging on photon energy by measuring the energy loss of the scattered protons in the forward detector TOTEM is an important feature. In this session J.Pinfold reported on photon-photon, photon-pomeron and double pomeron production at CDF. A recent workshop on photoproduction at collider energies at ECT*/Trento was devoted to UPC, the mini-proceedings can be found in~\cite{ect}. The reviews~\cite{ber,soff,bau,bns} and the most recent preprint~\cite{yr} reflect the gradual progess of the field. \section*{Acknowledgments} I would like to thank Frederic Kapusta for his kind invitation to this very pleasant and interesting conference at such a venerable place. \begin{footnotesize}
{ "timestamp": "2007-11-19T11:01:48", "yymm": "0711", "arxiv_id": "0711.2882", "language": "en", "url": "https://arxiv.org/abs/0711.2882" }
\section*{Acknowledgments} We thank P. Minnhagen and K. Schoutens for many enlightening discussions and for a critical reading of the manuscript. G. N. would like also to thank N. Kitanine for interesting discussions on a related lattice quantum integrable model \cite{PhM,Kolya}. G. N. is supported by the ANR programm MIB-05 JC05-52749. \section*{Appendix: reminder of TM for the JJL} Here we briefly summarize the main results of our theory, the TM, for the fully frustrated JJL \cite{noi3}\cite{noi4}. We first construct the bosonic theory and show that its energy momentum tensor fully reproduces the Hamiltonian of eq. (\ref{ha3}) for the JJL. That allows us to describe the JJL excitations in terms of the primary fields $V_{\alpha }\left( z\right) $% . Then we show that it is possible to construct the $N-$vertices correlator for the torus topology in $2D$ (basically by letting the edge to evolve in ``time''\ and to interact with external vertex operators placed at different points). We assume that a suitable correlator is apt to describe the ground state wave function of the JJL at $T=0$ temperature and then perform an analysis of the symmetry properties of its center of charge wave function (conformal blocks), which emerge in the presence of vortices carrying half quantum of flux ($\frac{1}{2}\left( \frac{hc}{2e}\right) $). Let us focus on the $m$-reduction procedure \cite{cgm4} for the special $m=2$ case (see Ref. \cite{cgm2} for the general case), since we are interested in a system with a $Z_{2}$ symmetry and choose the ``bosonic''\ theory \cite {noi3}\cite{noi4}, which well adapts to the description of a system with Cooper pairs of electric charge $2e$ in the presence of a topological defect \cite{noi1}, i.e. a fully frustrated JJL. To each of the two legs (edges) of the ladder we assign a chirality, so making a correspondence between up-down leg and left-right chirality states. Let us now write each phase field as the sum $\varphi ^{\left( a\right) }\left( x\right) =\varphi _{L}^{\left( a\right) }\left( x\right) +\varphi _{R}^{\left( a\right) }\left( x\right) $\ of left and right moving fields defined on the half-line because of the topological defect located in $x=0$. Then let us define for each leg the two chiral fields $\varphi _{e,o}^{\left( a\right) }\left( x\right) =\varphi _{L}^{\left( a\right) }\left( x\right) \pm \varphi _{R}^{\left( a\right) }\left( -x\right) $, each defined on the whole $x-$axis \cite{boso}. In such a framework the dual fields $\varphi _{o}^{\left( a\right) }\left( x\right) $\ are fully decoupled because the corresponding boundary interaction term in the Hamiltonian does not involve them \cite{affleck}; they are involved in the definition of the conjugate momenta $\Pi _{\left( a\right) }=\left( \partial _{x}\varphi _{o}^{\left( a\right) }\right) =\left( \frac{\partial }{\partial \varphi _{e}^{\left( a\right) }}\right) $\ present in the quantum Hamiltonian. Performing the change of variables $\varphi _{e}^{\left( 1\right) }=X+\phi $, $\varphi _{e}^{\left( 2\right) }=X-\phi $\ ($\varphi _{o}^{\left( 1\right) }=\overline{X}+\overline{\phi }$, $\varphi _{o}^{\left( 2\right) }=\overline{X}-\overline{\phi }$\ for the dual ones) we get the quantum Hamiltonian (\ref{ha3}) but now all the fields are chiral ones. Finally let us identify in the continuum such chiral phase fields $% \varphi _{e}^{\left( a\right) }$, $a=1,2$, each defined on the corresponding leg, with the two chiral fields $Q^{\left( a\right) }$, $a=1,2$\ of the TM with central charge $c=2$. As a result of the $2$-reduction procedure \cite{cgm2}\cite{cgm4} we get a $% c=2$ orbifold CFT, the TM, whose fields have well defined transformation properties under the discrete $Z_{2}$ (twist) group, which is a symmetry of the TM. Its primary fields content can be expressed in terms of a $Z_{2}$% -invariant scalar field $X(z)$, given by \begin{equation} X(z)=\frac{1}{2}\left( Q^{(1)}(z)+Q^{(2)}(z)\right) , \label{X} \end{equation} describing the continuous phase sector of the theory, and a twisted field \begin{equation} \phi (z)=\frac{1}{2}\left( Q^{(1)}(z)-Q^{(2)}(z)\right) , \label{phi} \end{equation} which satisfies the twisted boundary conditions $\phi (e^{i\pi }z)=-\phi (z)$ \cite{cgm2}. More explicitly such a field can be written in terms of the left and right moving components $\varphi _{L}^{\left( 1\right) }$, $\varphi _{R}^{\left( 2\right) }$ as we stated above; then the Mobius boundary conditions given in eq. (\ref{blr}) are described by the boundary conditions for $\phi $. This will be more evident for closed geometries, i.e. for the torus case, where the magnetic impurity gives rise to a line defect in the bulk, so allowing us to resort to the folding procedure and introduce boundary states \cite{noi1}\cite{noi2}. Such a procedure is used in the literature to map a problem with a defect line (as a bulk property) into a boundary one, where the defect line appears as a boundary state of a theory which is not anymore chiral and its fields are defined in a reduced region which is one half of the original one. Our approach, the TM, is a chiral description of that, where the chiral $\phi $\ field defined in ($-L/2$, $% L/2)$ describes both the left moving component and the right moving one defined in ($-L/2$, $\ 0$), ($0$, $L/2$) respectively, in the folded description \cite{noi1}\cite{noi2}. Furthermore to make a connection with the TM we consider more general gluing conditions: \[ \phi _{L}(x=0)=\mp \phi _{R}(x=0), \] the $-$($+$) sign staying for the twisted (untwisted) sector. We are then allowed to use the boundary states given in \cite{Affleck} for the $c=1$ orbifold at the Ising$^{2}$ radius. The $X$ field, which is even under the folding procedure, does not suffer any change in boundary conditions \cite {noi1} while condition (\ref{blr}) is naturally satisfied by the twisted field $\phi \left( z\right) $. So topological order can be discussed referring to the characters with the implicit relation to the different boundary states (BS) present in the system \cite{noi1}. These BS should be associated to different kinds of linear defects compatible with conformal invariance. The fields in eqs. (\ref{X})-(\ref{phi}) coincide with the ones introduced in eq. (\ref{ha3}). In fact the energy momentum tensor for such fields fully reproduces the second quantized Hamiltonian of eq. (\ref{ha3}). The whole TM theory decomposes into a tensor product of two CFTs, a twisted invariant one with $c=\frac{3}{2}$ and the remaining $c=\frac{1}{2}$ one realized by a Majorana fermion in the twisted sector. In the $c=\frac{3}{2}$ sub-theory the primary fields are composite vertex operators $V\left( z\right) =U_{X}^{\alpha _{l}}\left( z\right) \psi \left( z\right) $ or $V_{qh}\left( z\right) =U_{X}^{\alpha _{l}}\left( z\right) \sigma \left( z\right) $, where \begin{equation} U_{X}^{\alpha _{l}}\left( z\right) =\frac{1}{\sqrt{z}}:e^{i\alpha _{l}X(z)}: \label{char} \end{equation} is the vertex of the continuous\ sector with $\alpha _{l}=\frac{l}{2}$, $% l=1,...,4$ for the $SU(2)$ Cooper pairing symmetry used here. Regarding the other\ component, the highest weight state in the isospin sector, it can be classified by the two chiral operators: \begin{equation} \psi \left( z\right) =\frac{1}{2\sqrt{z}}\left( :e^{i\sqrt{2}\phi \left( z\right) }:+:e^{i\sqrt{2}\phi \left( -z\right) }:\right) ,~~~~~~\overline{% \psi }\left( z\right) =\frac{1}{2\sqrt{z}}\left( :e^{i\sqrt{2}\phi \left( z\right) }:-:e^{i\sqrt{2}\phi \left( -z\right) }:\right) ; \label{neu1} \end{equation} which correspond to two $c=\frac{1}{2}$ Majorana fermions with Ramond (invariant under the $Z_{2}$ twist) or Neveu-Schwartz ($Z_{2}$ twisted) boundary conditions \cite{cgm2}\cite{cgm4} in a fermionized version of the theory. The Ramond fields are the degrees of freedom which survive after the tunnelling and the parity symmetry, which exchanges the two Ising fermions, is broken. Besides the fields appearing in eq. (\ref{neu1}), there are the $% \sigma \left( z\right) $ fields, also called the twist fields, which appear in the quasi-hole primary fields $V_{qh}\left( z\right) $. The twist fields have non local properties and decide also for the non trivial properties of the vacuum state, which in fact can be twisted or not in our formalism. Starting from the primary fields $V_{\alpha }\left( z\right) $ we can now construct the non perturbative ground state wave function of the JJL system for the torus topology. It turns out that by construction it results as a coherent superposition of gaussian states with all the non trivial global properties of the order parameter.{\bf \ }In fact by using standard conformal field theory techniques it is now possible to generate the torus topology, starting from the edge theory, just defined above. That is realized by evaluating the $N$-vertices correlator \begin{equation} \left\langle n\right| V_{\alpha }\left( z_{1}\right) \ldots V_{\alpha }\left( z_{N}\right) e^{2\pi i\tau L_{0}}\left| n\right\rangle , \end{equation} where $V_{\alpha }\left( z_{i}\right) $ is the generic primary field representing the excitation at $z_{i}$, $L_{0}$ is the Virasoro generator for dilatations and $\tau $ the proper time. The neutrality condition $\sum \alpha =0$ must be satisfied and the sum over the complete set of states $% \left| n\right\rangle $ is indicating that a trace must be taken. It is very illuminating for the non expert reader to pictorially represent the above operation in terms of an edge state (that is a primary state defined at a given $\tau $) which propagates interacting with external fields at $% z_{1}\ldots z_{N}$ and finally getting back to itself. In such a way a $2D$ surface is generated with the torus topology. It is interesting to observe that such a procedure is equivalent to the coherent insertion of correlated relevant vortices (as provided by the CFT description) at positions $% z_{1}\ldots z_{N}$, as they appear in the non perturbative ground state of the physical JJL system.{\bf \ }From such a picture it is evident then how the degeneracy of the non perturbative ground state is closely related to the number of primary states. Furthermore, since in this letter we are interested in the understanding of the topological properties of the system, we can consider only the center of charge contribution in the above correlator, so neglecting its short distances properties. To such an extent the one-point functions are extensively reported in the following. On the torus \cite{cgm4} the TM primary fields are described in terms of the conformal blocks of the $Z_{2}$-invariant $c=\frac{3}{2}$ sub-theory and of the non invariant $c=\frac{1}{2}$ Ising model, so reflecting the decomposition on the plane above outlined. The characters $\bar{\chi}% _{0}(0|\tau )$, $\bar{\chi}_{\frac{1}{2}}(0|\tau )$, $\bar{\chi}_{\frac{1}{16% }}(0|\tau )$ express the primary fields content of the Ising model \cite{cft} with Neveu-Schwartz ($Z_{2}$ twisted) boundary conditions \cite{cgm4}, while \begin{eqnarray} \chi _{(0)}^{c=3/2}(0|w_{c}|\tau ) &=&\chi _{0}(0|\tau )K_{0}(w_{c}|\tau )+\chi _{\frac{1}{2}}(0|\tau )K_{2}(w_{c}|\tau )\,, \label{mr1} \\ \chi _{(1)}^{c=3/2}(0|w_{c}|\tau ) &=&\chi _{\frac{1}{16}}(0|\tau )\left( K_{1}(w_{c}|\tau )+K_{3}(w_{c}|\tau )\right) , \label{mr2} \\ \chi _{(2)}^{c=3/2}(0|w_{c}|\tau ) &=&\chi _{\frac{1}{2}}(0|\tau )K_{0}(w_{c}|\tau )+\chi _{0}(0|\tau )K_{2}(w_{c}|\tau ) \label{mr3} \end{eqnarray} represent those of the $Z_{2}$-invariant $c=\frac{3}{2}$ CFT. They are given in terms of a ``charged''\ $K_{\alpha }(w_{c}|\tau )$ contribution: \begin{equation} K_{2l+i}(w|\tau )=\frac{1}{\eta \left( \tau \right) }\;\Theta \left[ \begin{array}{c} \frac{2l+i}{4} \\[6pt] 0 \end{array} \right] (2w|4\tau )\,,\qquad \text{with }l=0,1\text{ and }i=0,1\,, \label{chp} \end{equation} and a ``isospin''\ one $\chi _{\beta }(0|\tau )$, (the conformal blocks of the Ising Model), where $w_{c}=\dfrac{1}{2\pi i}\,\ln z_{c}$ is the torus variable of the ``charged''\ component while the corresponding argument of the isospin block is $w_{n}=0$ everywhere. If we now turn to the whole $c=2$ theory, the characters of the twisted sector are given by: \begin{eqnarray} \chi _{(0)}^{+}(0|w_{c}|\tau ) &=&\bar{\chi}_{\frac{1}{16}}(0|\tau )\left( \chi _{0}+\chi _{\frac{1}{2}}\right) (0|\tau )\left( K_{0}+K_{2}\right) (w_{c}|\tau ), \label{tw1} \\ \chi _{(1)}^{+}(0|w_{c}|\tau ) &=&\chi _{\frac{1}{16}}(0|\tau )\left( \bar{% \chi}_{0}+\bar{\chi}_{\frac{1}{2}}\right) (0|\tau )\left( K_{1}+K_{3}\right) (w_{c}|\tau ), \label{tw2} \end{eqnarray} for the $A-P$ sector and by: \begin{eqnarray} \chi _{(0)}^{-}(0|w_{c}|\tau ) &=&\bar{\chi}_{\frac{1}{16}}(0|\tau )\left( \chi _{0}-\chi _{\frac{1}{2}}\right) (0|\tau )\left( K_{0}-K_{2}\right) (w_{c}|\tau ), \label{tw3.} \\ \chi _{(1)}^{-}(0|w_{c}|\tau ) &=&\chi _{\frac{1}{16}}(0|\tau )\left( \bar{% \chi}_{0}-\bar{\chi}_{\frac{1}{2}}\right) (0|\tau )\left( K_{1}+K_{3}\right) (w_{c}|\tau ), \label{tw4.} \end{eqnarray} for the $A-A$ one. Furthermore the characters of the untwisted sector are \cite{cgm4}: \begin{align} \tilde{\chi}_{(0)}^{-}(0|w_{c}|\tau )& =\left( \bar{\chi}_{0}\chi _{0}-\bar{% \chi}_{\frac{1}{2}}\chi _{\frac{1}{2}}\right) (0|\tau )K_{0}(w_{c}|\tau )+\left( \bar{\chi}_{0}\chi _{\frac{1}{2}}-\bar{\chi}_{\frac{1}{2}}\chi _{0}\right) (0|\tau )K_{2}\,(w_{c}|\tau ), \label{vac1.} \\ \tilde{\chi}_{(1)}^{-}(0|w_{c}|\tau )& =\left( \bar{\chi}_{0}\chi _{\frac{1}{% 2}}-\bar{\chi}_{\frac{1}{2}}\chi _{0}\right) (0|\tau )K_{0}(w_{c}|\tau )+\left( \bar{\chi}_{0}\chi _{0}-\bar{\chi}_{\frac{1}{2}}\chi _{\frac{1}{2}% }\right) (0|\tau )K_{2}\,(w_{c}|\tau ), \end{align} for the $P-A$ sector while for the $P-P$ sector we have: \begin{align} \tilde{\chi}_{\alpha }^{+}(0|w_{c}|\tau )& =\frac{1}{2}\left( \bar{\chi}_{0}-% \bar{\chi}_{\frac{1}{2}}\right) (0|\tau )\left( \chi _{0}-\chi _{\frac{1}{2}% }\right) (0|\tau )(K_{0}-K_{2})(w_{c}|\tau )\,, \\ \tilde{\chi}_{\beta }^{+}(0|w_{c}|\tau )& =\frac{1}{2}\left( \bar{\chi}_{0}+% \bar{\chi}_{\frac{1}{2}}\right) (0|\tau )\left( \chi _{0}+\chi _{\frac{1}{2}% }\right) (0|\tau )(K_{0}+K_{2})(w_{c}|\tau ), \label{vac4.} \end{align} and \begin{equation} \tilde{\chi}_{\gamma }^{+}(0|w_{c}|\tau )=\bar{\chi}_{\frac{1}{16}}(0|\tau )\chi _{\frac{1}{16}}(0|\tau )\left( K_{1}+K_{3}\right) (w_{c}|\tau ). \end{equation} Let us comment that the above factorization expresses the parity selection rule ($m$-ality), which gives a gluing condition for the ``charged''\ and ``isospin''\ excitations. It is worth underlining that in the $P-P$ sector, unlike for the other sectors, modular invariance constraint requires the presence of three different characters. The {\it isospin} operator content of the character $% \tilde{\chi}_{\gamma }^{+}(0|w_{c}|\tau )$ clearly evidences its peculiarity with respect to the other states of the periodic (even ladder) case. Indeed it is characterized by two twist fields ($\Delta =1/16$) in the {\it isospin} components. The occurrence of the {\it double} twist in the state described by $\tilde{\chi}_{\gamma }^{+}(0|w_{c}|\tau )$ is simply the reason why such a state is a periodic state. Indeed, being an {\it isospin} twist field the representation in the continuum limit of a magnetic impurity (a half flux quantum trapping\ or equivalently a kink), the double twist corresponds to a double half flux quantum trapping, i.e. one flux quantum, typical of the periodic configuration. The above analysis would suggest that the $P-P$ state described by $\tilde{% \chi}_{\gamma }^{+}(0|w_{c}|\tau )$ embeds in the continuum limit a kink-antikink excitation, i.e. it represents an excited state in the $P-P$ sector. In this way, as it happens for all the other sectors, the $P-P$ sector is left with just two degenerate ground states ( $\tilde{\chi}% _{\alpha }^{+}(0|w_{c}|\tau )$ and $\tilde{\chi}_{\beta }^{+}(0|w_{c}|\tau )$% ) and, as expected on a pure topological base, the ground state degeneracy in the torus topology is the double of that of the disk. Let us now present the full list of character transformations under the insertion of a magnetic flux quantum through the hole of the closed ladder. In the even closed JJ ladder configuration, we have that the two ground state wave functions of the $P-A$ sector decouple, being \begin{equation} {\cal T}_{1/2}\tilde{\chi}_{(0)}^{-}(0|w_{c}|\tau )=0\,,\text{ \ }{\cal T}% _{1/2}\tilde{\chi}_{(1)}^{-}(0|w_{c}|\tau )=0. \label{t(1/2)-PA} \end{equation} Concerning the $P-P$ sector, we have: \begin{equation} {\cal T}_{1/2}\tilde{\chi}_{\alpha }^{+}(0|w_{c}|\tau )=0 \label{t(1/2)-PPa} \end{equation} and \begin{equation} {\cal T}_{1/2}\tilde{\chi}_{\beta }^{+}(0|w_{c}|\tau )=\tilde{\chi}_{\gamma }^{+}(0|w_{c}|\tau )\,\text{ \ \ \ \ }(\ {\cal T}_{1/2}\tilde{\chi}_{\gamma }^{+}(0|w_{c}|\tau )=\tilde{\chi}_{\beta }^{+}(0|w_{c}|\tau )\,\text{\ }). \label{t(1/2)-PPbc} \end{equation} Such transformations show the instability of the $P-P$ sector under the insertion of a flux quantum through the hole of the closed ladder. More precisely the state $\tilde{\chi}_{\alpha }^{+}(0|w_{c}|\tau )$ decouples while the state $\tilde{\chi}_{\beta }^{+}(0|w_{c}|\tau )$ gets excited to the state with a kink-antikink configuration $\tilde{\chi}_{\gamma }^{+}(0|w_{c}|\tau )$. Furthermore in the odd closed JJ ladder configuration, we have that the two ground state wave functions of the $A-A$ sector decouple, being \begin{equation} {\cal T}_{1/2}\chi _{(0)}^{-}(0|w_{c}|\tau )=0\,,\text{ \ }{\cal T}% _{1/2}\chi _{(1)}^{-}(0|w_{c}|\tau )=0. \label{t(1/2)-AA} \end{equation} Concerning the $A-P$ sector, we have that the two ground state wave functions transform as: \begin{equation} {\cal T}_{1/2}\chi _{(0)}^{+}(0|w_{c}|\tau )=\chi _{(1)}^{+}(0|w_{c}|\tau )\,,\text{ \ }{\cal T}_{1/2}\chi _{(1)}^{+}(0|w_{c}|\tau )=\chi _{(0)}^{+}(0|w_{c}|\tau )\,. \label{t(1/2)-AP} \end{equation} Concluding, the full set of transformations, here presented, allows to claim the following simple and clear picture: {\it the odd closed JJL configuration is the only one which is stable under the insertion of a magnetic flux quantum through the central hole; moreover, in such odd JJL configuration such a magnetic flux insertion simply implements the flipping process between the two degenerate ground states }$\left| 0\right\rangle $% {\it \ and }$\left| 1\right\rangle ${\it .}
{ "timestamp": "2007-11-27T14:06:24", "yymm": "0711", "arxiv_id": "0711.4245", "language": "en", "url": "https://arxiv.org/abs/0711.4245" }
"\\section{Introduction}\r\n\r\nLet $G$ be a compact, $1$--connected and simple Lie group, namely, $(...TRUNCATED)
{"timestamp":"2010-08-31T02:01:14","yymm":"0711","arxiv_id":"0711.2541","language":"en","url":"https(...TRUNCATED)
"\\section{Introduction}\\label{s4.introduction}\n\nIn the first paper of this series \\citep[][here(...TRUNCATED)
{"timestamp":"2008-02-09T00:51:29","yymm":"0711","arxiv_id":"0711.3071","language":"en","url":"https(...TRUNCATED)
"\n\\section{Introduction}\n\n\n\nUnderstanding the dynamics of pion production in nucleon-nucleon c(...TRUNCATED)
{"timestamp":"2007-11-17T17:35:45","yymm":"0711","arxiv_id":"0711.2748","language":"en","url":"https(...TRUNCATED)
"\\section{Introduction}\n\nIn the present work we deal with pattern avoidance on words. This\ntopic(...TRUNCATED)
{"timestamp":"2007-11-21T15:53:22","yymm":"0711","arxiv_id":"0711.3387","language":"en","url":"https(...TRUNCATED)
End of preview.