1: \documentclass[final]{aipproc}
2: \layoutstyle{6x9}
3:
4:
5:
6:
7: \newcommand{\be}{\begin{equation}}
8: \newcommand{\ee}{\end{equation}}
9: \newcommand{\hmp}{h^{-1}Mpc}
10: \newcommand{\bef}{\begin{figure}}
11: \newcommand{\eef}{\end{figure}}
12:
13: \newcommand{\etal}{et al.}
14: \def\eg{{e.g.}}
15: \def\ie{{i.e.}}
16: \newcommand{\kms}{\,{\rm km}\;{\rm s}^{-1}}
17: \newcommand{\hubunits}{\,\kms\;{\rm Mpc}^{-1}}
18: \newcommand{\hmpc}{\,h^{-1}\;{\rm Mpc}}
19: \newcommand{\hkpc}{\,h^{-1}\;{\rm kpc}}
20: \newcommand{\msun}{M_\odot}
21: \newcommand{\K}{\,{\rm K}}
22: \newcommand{\cm}{{\rm cm}}
23: \newcommand{\bea}{\begin{eqnarray}}
24: \newcommand{\eea}{\end{eqnarray}}
25: \newcommand{\cd}{{\langle n(r) \rangle_p}}
26: \newcommand{\Mpc}{{\rm Mpc}}
27: \newcommand{\kpc}{{\rm kpc}}
28: \newcommand{\xir}{{\xi(r)}}
29: \newcommand{\xrp}{{\xi(r_p,\pi)}}
30: \newcommand{\xsirpi}{{\xi(r_p,\pi)}}
31: \newcommand{\wrp}{{w_p(r_p)}}
32: %\newcommand{\gr}{{^{0.1}g-r}}
33: \newcommand{\gr}{{g-r}}
34: \newcommand{\Navg}{N_{\rm avg}}
35: \newcommand{\Mmin}{M_{\rm min}}
36: \newcommand{\fiso}{f_{\rm iso}}
37: \newcommand{\Mr}{M_r}
38: \newcommand{\rp}{r_p}
39: \newcommand{\zmax}{z_{\rm max}}
40: \newcommand{\zmin}{z_{\rm min}}
41:
42:
43: \def\spose#1{\hbox to 0pt{#1\hss}}
44: \def\ltapprox{\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$}}
45: \raise 2.0pt\hbox{$\mathchar"13C$}}}
46: \def\gtapprox{\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$}}
47: \raise 2.0pt\hbox{$\mathchar"13E$}}}
48: \def\inapprox{\mathrel{\spose{\lower 3pt\hbox{$\mathchar"218$}}
49: \raise 2.0pt\hbox{$\mathchar"232$}}}
50:
51:
52: \begin{document}
53:
54: \title{Gravitational clustering: an overview}
55:
56: \classification{05.40.-a, 95.30.Sf}
57:
58: \keywords{Gravitation, structure formation, cosmology}
59:
60: \author{Francesco Sylos Labini}{
61: address={``E. Fermi'' Center, Via Panisperna 89 A, Compendio del
62: Viminale, I-00184 Rome, Italy,\\ \& ISC-CNR, Via dei Taurini 19,
63: I-00185 Rome, Italy.}
64: }
65:
66:
67: \begin{abstract}
68: We discuss the differences and analogies of gravitational clustering
69: in finite and infinite systems. The process of collective, or violent,
70: relaxation leading to the formation of quasi-stationary states is one
71: of the distinguished features in the dynamics of self-gravitating
72: systems. This occurs, in different conditions, both in a finite than
73: in an infinite system, the latter embedded in a static or in an
74: expanding background. We then discuss, by considering some simple and
75: paradigmatic examples, the problems related to the definition of a
76: mean-field approach to gravitational clustering, focusing on role of
77: discrete fluctuations. The effect of these fluctuations is a basic
78: issue to be clarified to establish the range of scales and times in
79: which a collision-less approximation may describe the evolution of a
80: self-gravitating system and for the theoretical modeling of the
81: non-linear phase.
82: \end{abstract}
83:
84: \maketitle
85:
86: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
87: %% MAINMATTER
88: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
89:
90:
91: \section{Introduction}
92:
93:
94: As discussed in various papers in this volume (see
95: e.g.\cite{intro,campa}) equilibrium properties of long-range interacting
96: systems require a non-trivial analysis as standard thermodynamics
97: techniques do not simply apply when dealing with pair-interactions
98: decaying with sufficiently small exponents. Many interesting and
99: unsolved problems lie in the out-of-equilibrium dynamics of systems
100: with long-range interaction about which very little is known from a
101: theoretical point of view (see also
102: \cite{chavanis} in this volume).
103: The understanding of the thermodynamics and dynamics of systems of
104: particles interacting only through their mutual Newtonian self-gravity
105: is of fundamental importance in cosmology and astrophysics. It
106: encompasses the range of physical scales relevant to the formation of
107: the largest structures in the Universe, down to those relevant to
108: stellar dynamics. The statistical mechanics of systems dominated by
109: gravity has been studied and applied in many different contexts in
110: astrophysics and cosmology (see
111: e.g.
112: \cite{lyndebell,pad_physrep,chavanis,chandra,chandra_revmodphy,
113: pad_book,pad_dtslri,saslaw1,saslaw2,binney,peebles,thierry}):
114: for example in the studies of globular clusters, galaxies and the
115: clustering in the expanding universe.
116:
117:
118: While systems with short range interactions can be usually studied
119: through laboratory experiments, gravitational systems can only be
120: observed in astrophysical contexts. Alternatively one may set up
121: numerical experiments which then represent the unique instrument to
122: study the dynamics of gravitational clustering. In this respect the
123: astrophysicist's perspective is usually to model some intricate
124: realistic systems, such as stellar or galaxy systems, having the aim
125: of understanding a specific set of observations. For example in the
126: cosmological context one uses very complicated initial conditions
127: (described by a large number of parameters) and needs a certain number
128: of important assumptions, from the way the universe expands to the
129: amount and type of dark matter which dominates the dynamics on the
130: relevant scales. This is so because, by studying gravitational
131: clustering, one would like to understand the relations between some
132: important observations of the cosmos. For example the studies of the
133: cosmic microwave background radiations provide with the information
134: about the initial conditions of the matter density field. The large
135: scale geometrical properties of the universe are deduced, for example,
136: through the measurements of the supernovae magnitude-redshift
137: relation. Galaxy redshift surveys map the present-day matter
138: distribution. The estimations of the mass-to-light ratio of
139: astrophysical objects is ultimately related to the abundance of dark
140: matter. The task of the model of cosmological structure formation is
141: thus to build a unified and coherent picture to explain these (and
142: other) observations of the universe at the largest scales
143: \cite{peacock}.
144:
145:
146: In statistical physics the problem of the evolution of
147: self-gravitating classical bodies has been relatively neglected,
148: primarily because of the intrinsic difficulties associated with the
149: attractive long-range nature of gravity and its singular behavior at
150: vanishing separation. When approaching the problem of gravitational
151: clustering in the context of statistical mechanics it is natural to
152: start by reducing as much as possible the complexity of the analogous
153: cosmological or astrophysical problem. For example, in order to focus
154: on the essential aspects of the problem one may study gravitational
155: clustering without the expansion of the universe, and starting from
156: particularly simple initial conditions. With respect to the motivation
157: from cosmology/astrophysics, there is of course a risk: in simplifying
158: we may loose some essential elements which change the nature of
159: gravitational clustering. Even it were, it seems unlikely that we will
160: not learn something about the more complex cosmological/astrophysical
161: situations in addressing slightly different and simplified problems.
162:
163:
164:
165: A fundamental distinction has to be made between finite and infinite
166: systems. They have in common that the gravitational force on a
167: arbitrary point has contributions coming from all scales in the
168: system. However they differ for the fact that in the case of the
169: finite system there is a mean field force generated by the system as a
170: whole, which is related to its internal symmetries (e.g., spherical
171: symmetry) and which eventually will give rise to a global collapse of
172: the entire system. In an infinite space, in which the initial
173: fluctuations are non-zero and finite at all scales, the collapse of
174: larger and larger scales will continue ad infinitum: being no
175: geometric center there will not be a global collapse of the entire
176: system. The mean field dynamics is driven by system's fluctuations
177: which determine the gravitational force at different spatial scales.
178: The collapse occurring on larger and larger scales will clearly happen
179: at different times but it is characterized by the unique time-scale in
180: the system that is
181: %
182: \be
183: \label{tau}
184: \tau \sim \sqrt{G\rho_0}^{-1} \,.
185: \ee
186: %
187: This is in general the typical characteristic time scale of any
188: (finite or infinite) gravitational system with average mass density
189: $\rho_0$. For instance, as we discuss below, this is the time scale
190: predicted by the self-gravitating fluid approximation.
191:
192: The infinite system can therefore never reach a time independent
193: state, and in particular it will never reach a thermodynamic
194: equilibrium. Although so different, the finite and the infinite
195: systems share some subtle and important analogies which we briefly
196: discuss in what follows. We firstly discuss the general difficulties
197: related to the long-range character of gravity and the usual way to
198: make a mean-field approximation. Then we consider two basic examples
199: of a finite and of an infinite system: the simplest example of a
200: finite system is represented by an initially isolated spherical
201: distribution of randomly distributed points (i.e. a Poisson
202: distribution) --- this will be also discussed in the contribution by
203: Morikawa in this book \cite{morikawa}. Analogously the simplest
204: example of an infinite system is a Poisson distribution in an infinite
205: space. For this second case one may consider that the space background
206: is static (as Joyce in his contribution in this book \cite{joyce}) or
207: it is expanding (as Saslaw in his contribution \cite{saslaw}). We will
208: discuss these two cases briefly, outlining the analogies and the
209: differences. In the conclusions we try to point out which are the main
210: problems of the field.
211:
212:
213:
214:
215:
216: \section{Gravitational relaxation and mean field models}
217: \label{force}
218:
219:
220: The typical feature of short-range interactions system is that the
221: force between two particles (e.g. molecules of an ordinary gas) is
222: strong when they are very close to each other, i.e. when they repeal
223: each other strongly, and on large distances no force is exerted
224: between two particles. In this way particles most of time move at
225: nearly constant velocity, and then they are subject to violent and
226: short lived accelerations when they collide with one another. In this
227: situation, the typical mechanism for relaxation is that due to
228: two-body collisions. For self-gravitating particles the situation is
229: different because gravity is divergent at vanishing separation and it
230: is long range. Because of the former property the actual force on a
231: given particle maybe more or less strongly influenced by fluctuations
232: in its local neighborhood, while the latter property implies that a
233: particle is coupled with all other particles at all scales in the
234: system. Thus the studies of a self-gravitating systems have to
235: consider that many different range of scales, in principle all, maybe
236: almost equally important for the analysis of the dynamical properties.
237:
238:
239:
240: For this reason, one the principal problems of stellar dynamics
241: \cite{chandra,chandra_revmodphy} is concerned with the analysis of the nature
242: of the gravitational force acting, for example, on a star which is
243: member of a stellar system. In a general way we may broadly
244: distinguish between the influence of the system as a whole
245: $\vec{F}_{sm}$ and of the immediate neighborhoods $\vec{F}_{f}$. The
246: former will be smoothly varying function of position and time while
247: the latter will be subject to relatively rapid fluctuations:
248: \[
249: \vec{F} = \vec{F}_{sm} + \vec{F}_{f} \;.
250: \]
251: The second term describes the fluctuations that are related to the
252: underlying statistical properties of the particle distribution: their
253: effects can be evaluated, under certain assumptions, in a stochastic
254: sense.
255:
256: The problem is general is to establish a criterion to quantitatively
257: determine the circumstances under which
258: \be
259: \label{fapp}
260: \vec{F} \approx \vec{F}_{sm} \;,
261: \ee
262: and thus $\vec{F}_{f} \approx 0$, i.e. under which the system can be
263: treated by the mean field approximation. This consists in neglecting
264: the local interaction and to consider the global gravitational field
265: of the system. This is not an easy task for the general case. For
266: example in the simplest situation of an isolated self-gravitating
267: system in virial equilibrium it is possible to estimate a criterion
268: for using the approximation provided with Eq.\ref{fapp}, by considering
269: the role of two-body scattering
270: \cite{chandra,binney,saslaw2}. Whether this can be used for the more
271: general case of an out-of-equilibrium system is an open question
272: \cite{binneyknebe,diemond,power02}.
273:
274:
275:
276: As mentioned above, the mean field approximation consists in neglecting
277: the physics on small scales in the system and for this reason it
278: usually describes a collision-less system. We briefly recall the main
279: steps to obtain such an
280: approximation~\cite{binney,pad_physrep,saslaw2}. Let us divide the
281: phase space of a given system into a large number of cells in such a
282: way that (i) each cell is large enough to contain a macroscopic number
283: of particles (ii) each cell is small enough so that all particles in
284: the cell can be assumed to have the same average characteristic
285: properties of the cell. Thus the size of the cell should be large
286: enough to satisfy (i) but not so large to violate (ii). If such an
287: intermediate size can be defined at a generic time $t$ it is possible
288: to define smooth functions, as for example the one-point distribution
289: function. As the collapse will involve progressively larger and larger
290: scales this approximation may break down at a certain time, when
291: non-linear objects will be formed on the scale of the cell.
292:
293:
294:
295: In the approximations above discussed the one-point distribution
296: function will satisfy the Boltzmann equation. By neglecting the
297: collision terms one may obtain the collision-less Boltzmann equation
298: or Vlasov equation which is the basic tool to study the evolution of
299: self-gravitating systems. From the Vlasov-Poisson system of equations
300: one may derive, by considering other suitable approximations, the
301: equations describing the evolution of a self-gravitating fluid. In
302: this treatment the discrete nature of a particle distribution is
303: neglected in the dynamical evolution: we will come back on this point
304: below.
305:
306:
307:
308:
309:
310:
311: \section{Finite systems}
312: \label{finite}
313:
314: We now consider a very simple model to describe the dynamics of an
315: isolated finite system. Let us suppose to have an initial uniform
316: distribution of $N$ points of identical mass $m$ in a spherical volume
317: of radius $R_0$ and density
318: \[
319: \rho_0 = Nm \frac{3}{4\pi R_0^3} \;.
320: \]
321: We suppose that this system is isolated in an infinite space. Let us
322: consider the limit of a perfect uniform system
323: \[
324: \lim_{N \rightarrow \infty, \;\; m \rightarrow 0} Nm = \mbox{const} \;,
325: \]
326: where the fluctuations in the system also go to zero in the same limit.
327: The mass contained in a sphere of radius $r_0 \le R_0$ is just
328: \be
329: M(r_0) = \int_0^{r_0} 4\pi r^2 \rho_0 dr = \frac{4 \pi}{3} r_0^3 \rho_0 \;.
330: \ee
331: For reasons of spherical symmetry,
332: the force exerted on a particle of unitary mass in the point at distance
333: $r_0$ from the center of the sphere is due to the matter inside the
334: shell
335: \be
336: F(r_0)= -\frac{G M(r_0)}{r_0^2} \;.
337: \ee
338: This force is radial and directed toward the center of the sphere. For
339: simplicity we suppose that the initial velocities are zero so that the
340: initial total energy is equal to the initial potential energy and thus
341: it is negative. In addition, we make the assumption that different
342: shells do not overlap during the collapse. Thus the point initially
343: at $r_0$ will be attracted by the same mass $M(r_0)$ at all times but
344: with varying force. The equation of motion can be written as
345: \be
346: \label{scm1}
347: \ddot r(t) = -\frac{G M(r_0)}{r^2(t)} \;.
348: \ee
349: This can be integrated to give
350: \be
351: \left(\frac{dr}{dt}\right)^2 = 2GM(r_0) \left(
352: \frac{1} {r} - \frac{1} {r_0} \right) \;,
353: \ee
354: where $G$ is the Newton's constant.
355: The solution of this equation can be written in a parametric form
356: \bea
357: &&
358: r(\xi) = \frac{r_0}{2} (1 + \cos(\xi))
359: \\ \nonumber
360: &&
361: t(\xi) = \sqrt{\frac{3}{32 \pi G \rho_0}} \left( \xi + \sin(\xi) \right) \;,
362: \label{scm2}
363: \eea
364: where
365: \bea
366: &&
367: r(\xi=0)=r_0 \;\;\; \mbox{at} \;\;\; t(\xi=0)=0
368: \\ \nonumber
369: && r(\xi=\pi)=0 \;\;\;\mbox{at}
370: \; \; t(\xi=\pi)= \sqrt{\frac{3\pi}{32 G \rho_0}}
371: \equiv \tau_{scm} \;.
372: \eea
373: Thus this evolution describes the collapse from the time $t=0$ at
374: which the system has radius $R_0$ to the time $\tau_{scm}$ at which
375: the system has zero radius. At longer times the system will re-expand
376: up to reach the initial configurations, and then will continue these
377: oscillations periodically. It is interesting to note that in this
378: model the system never reaches a virialized state (although the total
379: energy is negative).
380:
381:
382: There are some essential aspects of the problem which this
383: approach is missing. It is in fact well known through the studies of
384: computer experiments, that a system of particles is well described by
385: Eq.\ref{scm2} at times prior to $\tau_{scm}$, while it differs for $t
386: \rightarrow \tau_{scm}$, when the particle system goes through a phase
387: of collective relaxation which brings a part of particles to form a
388: quasi-equilibrium configuration in virial equilibrium. Total collapse
389: will never occur because small initial inhomogeneities in the system,
390: always present in any particle distribution, will generate random
391: motions which will eventually stop the collapse (see Fig.\ref{fig1}).
392:
393: \begin{figure}
394: \includegraphics*[height=.3\textheight]{fig1.ps} \caption{Different
395: phases of the spherical cold collapse (projection of the $x-y$ plane
396: of a thin slice around the center of the system, chosen, at
397: different times, to have the approximately the same number of points
398: inside): top-left initial distribution, top-right at $t=0.8
399: \tau_{scm}$, bottom-left at $t=1 \tau_{scm}$ and bottom-right at
400: $t=2 \tau_{scm}$. From this latter time on the structure is
401: virialized and a forms quasi-stationary state.}
402: \label{fig1}
403: \end{figure}
404:
405:
406: In this situation the system forms a state characterized by a dense
407: central region in virial equilibrium surrounded by a low-density halo
408: made of particles with positive total energy. The subsequent evolution
409: is driven by collisions: particles in the high density central region
410: can undergo to two-body scatterings. The integrated effect of these
411: collisions is that a part of the particles gain some kinetic energy
412: while the others end up in a more bounded state. The long term
413: consequence of these close encounters is the so-called evaporation,
414: i.e. that the core becomes more and more bounded and looses little by
415: little its particles by ejecting them. Thus the asymptotic state
416: should be made of very few (in principle two) particles orbiting one
417: around the others, very close so to keep all the potential energy of
418: the system, and an ideal gas made by the other particles which can
419: move freely bringing the main part of the kinetic energy of the
420: system (see Fig.\ref{fig2}).
421:
422:
423:
424:
425: Lynden-Bell \cite{lyndebell}, who named the collective relaxation
426: process as ``violent relaxation'', made the first attempt to construct
427: a theory describing the process of gravitational collapse by
428: considering a collision-less system. While the process of violent
429: relaxation is the main mechanism for relaxation of self-gravitating
430: systems, in very different situations, it is still not completely
431: understood in its full details and the theory of Lynden-Bell has shown
432: various disagreements when compared with the results with N-body
433: simulations even for the simplest case illustrated above (the same
434: occurs with other theoretical attempts --- see e.g. \cite{violrelax}).
435:
436:
437: The violent relaxation process acts on a time scale $\tau_{scm}$ which
438: is much shorter than, for example, the typical time scale for two-body
439: collisions $\tau_2$, which sets the time scale for the evaporation of
440: the core-halo structure mentioned above. Indeed one may show
441: \cite{saslaw2,binney} that for a
442: system of $N$ particles this goes as
443: \[
444: \tau_2 \propto \frac{N}{\log(N)} \tau_{scm} \;.
445: \]
446: This situation outlines an important difference between systems with
447: long and short range interactions. As mentioned, for systems with
448: short-range forces the relaxation proceeds through
449: collisions. Self-gravitating systems instead relax through processes
450: other than the collisional relaxation, such as violent relaxation,
451: which operate on time scales which are much shorter than the two-body
452: relaxation time. In fact, the collisional relaxation time for several
453: astrophysical systems is larger than the age of the universe and for
454: many years this was considered a paradox (see e.g. \cite{saslaw2}):
455: how can globular cluster be in virial equilibrium if their relaxation
456: time (supposed to be given by $\tau_2$) is longer than the age of the
457: universe ? The answer was simply that the relaxation mechanism was
458: different from two-body collisions.
459:
460:
461: \begin{figure}
462: \includegraphics*[height=.3\textheight]{fig2.eps}
463: \caption{ Virial
464: ratio $b=2K/W$, where $K$ is the total kinetic energy of the bound
465: particles and $W$ their gravitational potential energy, as a function
466: of time: At times $t>\tau_{scm}$ we have that $b\approx -1$ and hence the
467: virial theorem is satisfied. }
468: \label{fig2}
469: \end{figure}
470:
471: Thus the complete theoretical characterization of the simple spherical
472: collapse of cold particles described above, is an important and still
473: open problem not only for astrophysics but also for cosmology. Indeed
474: in both contexts one observes, in numerical simulations, the
475: formations of core-halo structures (we will come back below on the
476: cosmological simulations and on the structures formed therein). Thus a
477: mechanism similar to the violent relaxation of a finite system is
478: essential also for the virialization of larger and larger clusters of
479: particles in an infinite system although with some important
480: differences
481: \cite{white,morikawa}. As in the finite system, also in the infinite
482: system the two-body relaxation time scale is much longer than the real
483: relaxation time of structures. Because the mean field force is due to
484: fluctuations, the time-scales for the collapse of an over-density of
485: given size is much longer in the infinite system than in the finite
486: system case.
487:
488:
489: The ultimate task of a theory describing the process of violent
490: relaxation should be the prediction of the quasi-stationary
491: (virialized) state, which can be generally described by the Vlasov
492: equation, given the initial conditions in terms of the phase space
493: density (see e.g. \cite{hansen}). In particular there is a full set of
494: properties of the virialized state, such as the density profile, the
495: velocity distribution and profile, and their relations, which seems to
496: be universal, i.e. appearing from many different initial conditions
497: (see e.g. \cite{morikawa}) and that are currently unexplained from a
498: theoretical point of view.
499:
500:
501: Finally we note that being the two-body relaxation time larger than
502: $\tau_{scm}$ when the number of particles is large enough, the
503: collisional effects can usually be neglected for self-gravitating
504: quasi-equilibrium virial states, which can be treated with an
505: appropriate non-collisional (i.e. Vlasov) approximation. However in
506: general the problem consists in the comprehension of the role of
507: collisions and of the terms related to discreteness (i.e. random
508: forces, etc.) during the out-of-equilibrium phase, i.e. during
509: collapse giving rise to a stationary configuration. We will come back
510: to this point in what follows.
511:
512:
513:
514: \section{Infinite systems}
515: \label{infinite}
516: \begin{figure}
517: \includegraphics[height=.3\textheight]{fig5.ps}
518: \caption{Evolution of the fluctuations and formation of structures
519: in a simulation (with periodic boundary conditions, representing the
520: infinite system case) started from cold Poisson initial conditions.
521: Structures form firstly on small scales (top-left and top-right) and
522: then propagate to larger and large scales (bottom-left and
523: bottom-right).}
524: \label{fig5}
525: \end{figure}
526: The problem of the evolution of self-gravitating classical bodies,
527: initially distributed very uniformly in infinite space, is as old as
528: Newton. Modern cosmology poses essentially the same problem as the
529: matter in the universe is now believed to consist predominantly of
530: almost purely self-gravitating particles --- so called dark matter ---
531: which is, at early times, indeed very close to uniformly distributed
532: in the universe, and at densities at which quantum effects are
533: completely negligible. Despite the age of the problem and the
534: impressive advances of modern cosmology in recent years, our
535: understanding of it remains, however, very incomplete. In its
536: essentials it is a simple well posed problem of classical statistical
537: mechanics.
538:
539: The standard paradigm for formation of large scale structure in the
540: universe is based on the growth of small initial density fluctuations
541: in a homogeneous and isotropic medium. In the currently most popular
542: cosmological models, a dominant fraction (more than 80 \%) of the
543: clustering matter in the universe is assumed to be in the form of
544: microscopic particles which interact essentially only by their
545: self-gravity. At the macroscopic scales of interest in cosmology the
546: evolution of the distribution of this matter is then very well
547: described by the Vlasov equation coupled with the Poisson equation
548: \cite{binney,pad_physrep,saslaw1,saslaw2}.
549:
550: A full solution, either analytical or numerical, of these equations
551: starting from appropriate initial conditions is not feasible. In
552: cosmology perturbative approaches to the problem, which treat the very
553: limited range of low to modest amplitude deviations from uniformity,
554: have been developed but numerical simulations are essentially the only
555: instrument beyond this regime and to study the non-linear phase
556: \footnote{By non-linear objects we mean the structures formed having a
557: typical density $\rho$ such that the density contrast is $\delta =
558: (\rho -\rho_0)/\rho_0 \gg 1$, where $\rho_0$ is the average density of
559: the distribution on large enough scales. By non-linear clustering we
560: mean the dynamical evolution leading to the formation of non-linear
561: objects.}. N-body simulations solve numerically for the evolution of
562: a system of $N$ particles interacting purely through gravity, with a
563: softening at very small scales\footnote{One of the main concerns is
564: the independency on the softening length of the relevant statistical
565: quantities}. The number of particles $N$ in the very largest current
566: simulations is $\sim 10^{10}$ \cite{millenium}, many more than two
567: decades ago, but still many orders of magnitude fewer than the number
568: of real dark matter particles ($\sim 10^{80}$ in a comparable volume
569: for a typical candidate). While such simulations constitute a very
570: powerful and essential tool, they lack the valuable guidance which a
571: fuller analytic understanding of the problem would provide. The
572: question inevitably arises of the accuracy with which these
573: ``macro-particles'' trace the desired correlation properties of the
574: theoretical models. This is the problem of discreteness in
575: cosmological N-body simulations. It is an issue which is of
576: considerable importance as cosmology requires ever more precise
577: predictions for its models for comparison with observations.
578:
579: As already mentioned, the infinite system can never reach a time
580: independent state (see Fig.\ref{fig5})\footnote{To simulate an
581: infinite system one considers a finite volume and infinite replicas
582: of it. The force acting on a point is due to all particles inside
583: the volume and all replicas, i.e. periodic boundary conditions are
584: used.}. However one of the important results from numerical
585: simulations of such systems in the context of cosmology is that the
586: system nevertheless reaches a kind of scaling regime, in which the
587: temporal evolution is equivalent to a rescaling of the spatial
588: variables \cite{efts88,smith2003}. This spatio-temporal scaling
589: relation is referred to as ``self-similarity''\footnote{Note that in
590: this context the term ``self-similarity'' is used in a completely
591: different meaning than usually in statistical physics. For this
592: reason we always use quotes to mark this fact.} and this behavior is
593: usually observed for the time evolution of the two-point correlation
594: function or its Fourier transform, the power spectrum (see
595: Fig.\ref{fig4}). Before describing this evolution in more detail let
596: us come back to the determination of the gravitational force on an
597: average particle in a certain distribution and to the approximations
598: usually employed to study clustering in an infinite system.
599: %
600: \begin{figure}
601: \includegraphics[height=.3\textheight]{fig4.eps} \caption{Evolution
602: of the two-point correlation function in a Poisson simulation with periodic
603: boundary conditions (representing an infinite system). The temporal
604: evolution of this statistical quantity is equivalent to a rescaling
605: of the spatial variables. The time indicated in the labels is given
606: in units of the dynamical time scale $\tau=\sqrt{4 \pi G
607: \rho_0}^{-1}$ of the system. The distance scale $\Lambda_i$ marks
608: the initial average distance between nearest particles.}
609: \label{fig4}
610: \end{figure}
611: %
612:
613:
614:
615: \subsection{The force distribution}
616:
617: Let us consider a uniform particle distribution
618: \cite{book} in a finite volume $V$.
619: The actual value of the gravitational force per unit mass acting on a
620: fixed ``test'' particle belonging to the system, supposed to be on the
621: point $\vec{x}$ due to all the other $N$ system particles is given
622: by
623: \be
624: \label{force1}
625: \vec{F}(\vec{x}) = - Gm \sum_{i=1}^N \frac{ \vec{x} - \vec{x}_i}
626: {|\vec{x} -\vec{x}_i|^3}
627: \ee
628: where $m$ is the particle's mass, supposed to be the same for all
629: particles, and $\vec{x}_i$ the position vector of the $i^{th}$
630: particle. The sum in Eq.\ref{force1} is extended to all the particles
631: other than the test one, which are contained in the system volume $V$
632: Eventually $V \rightarrow \infty$ and $N \rightarrow \infty$ in such a
633: way that the average number density $n_0=N/V$ remains constant. The
634: actual value of $\vec{F}$ clearly depends on the nature of the local
635: particle distribution, and hence, it will be in general subjected to
636: stochastic fluctuations. These fluctuations determine a more or less
637: broad force probability density function (PDF) $P(\vec{F})$. Because
638: of statistical isotropy the direction of $\vec{F}$ is completely
639: random with equal probability in each direction, i.e. $P(\vec{F})$
640: depends only on $F=|\vec{F}|$. Thus one may simply consider
641: $P(\vec{F})d
642: \vec{F} =4\pi F^2 W(F)dF$, where $W(F)$ is the PDF of the modulus of
643: the force.
644: For reasons of spherical symmetry the average
645: gravitational force acting on one system's particle and due to the
646: rest of the system vanishes. Any force acting on a particle is due to
647: fluctuations away from exact (i.e. deterministic) spherical symmetry.
648:
649:
650: Chandrasekhar \cite{chandra_revmodphy} (see also \cite{book,force})
651: has considered the behavior of the PDF of the Newtonian gravitational
652: force arising from a statistically isotropic
653: (infinite) Poisson distribution of sources. He showed that applying
654: the Markov method, it is possible to compute exactly the PDF, known as
655: the Holtzmark distribution, of the gravitational force acting on a
656: test particle in the system.
657:
658: A very simple and approximate way to compute this PDF is to consider
659: only the contribution of the first neighbor: in this way from the
660: nearest neighbor (NN) probability distribution one may get the PDF of
661: the force. An interesting point is that the Holtzmark distribution and
662: the approximated PDF derived by considering the contribution of only
663: the NN agree very well in the large $F$ region (see Fig.\ref{fig3}).
664: %
665: \begin{figure}
666: \includegraphics[height=.3\textheight]{fig3.eps} \caption{Holtzmark
667: distribution and the PDF inferred if only NNs contribute, i.e.,
668: $W_{NN}(F)dF=\omega (r) dr$ being $\omega(r)$ the NN PDF. The
669: agreement is very good in the strong field limit where $W(F) \sim
670: F^{-5/2}$. At weak field the PDF due to the NN has a sharp cut-off
671: while the full Holtzmark distribution shows a more gentle decrease
672: (see discussion in \cite{book,force,earlytimes}).}
673: \label{fig3}
674: \end{figure}
675: %
676: The region where they
677: differ most is when $F \rightarrow 0$. This is due to the fact that a
678: weak force can arise only from a more or less symmetric configuration
679: of particles around the test one in which fluctuations are determined
680: by many particle effects, and hence the NN approximation
681: fails. Instead in the strong field limit we may almost neglect the
682: contribution to the force from far away points, because the main
683: contribution is due to the limit $r \rightarrow 0$ in the elementary
684: interaction (i.e. it comes from the NN). Note that, as in the NN
685: case, because of the behavior of $W(F)$ for $F\rightarrow \infty$,
686: $\langle F^2\rangle$ diverges. This is due to the singularity of the
687: particle-particle gravitational interaction $\sim 1/r^2$ at $r=0$
688: together with the fact that in the Poisson distribution there is no
689: explicit positive minimal distance between particles (i.e. no lower
690: cut-off), as they are permitted to be at an arbitrarily small distance
691: from one another.
692:
693: The extension of the Holtzmark distribution to other statistically
694: homogeneous and isotropic distributions is in general not
695: straightforward, the main complication being introduced by the presence
696: of spatial correlations among particles. One may find in
697: \cite{book,force,masucci,pellegrini} (and references therein) the
698: derivation of the PDF, under certain assumptions, for particle
699: distributions with non-trivial correlations.
700:
701:
702:
703:
704: \subsection{The infinite static case}
705:
706: In the previous derivation we used an important assumption (see also
707: \cite{joyce} in this volume), which we now examine in more detail.
708: The sum defined in Eq.\ref{force1} is not well defined because it is
709: only conditionally convergent for $V \rightarrow \infty$, i.e. its
710: result depends on the order in which the single terms are summed.
711: Thus the assumption is that the sum has been taken symmetrically with
712: respect to the point $\vec{x}$. Other prescriptions will in general
713: lead to different results. To see more clearly this fact let us
714: consider the local particle density embedded in a uniform background
715: of negative mass density so that the microscopic mass density is
716: \be
717: \delta \rho(\vec{x}) = m(n(\vec{x}) -n_0)
718: \ee
719: where $n(\vec{x})$ is the local particle density with average density
720: $n_0$ and the second term represents the uniform background. In this
721: situation the force on a particle is a well defined stochastic
722: quantity, i.e. it does not depend on how the volume $V$ is sent to
723: infinity \cite{force}. Given that the limit $V \rightarrow \infty$
724: does not depend on the way the limit is performed, one can choose for
725: simplicity to take the volume $V$ symmetric with respect to the point
726: $\vec{x}$ where the force is computed. In such a volume the
727: contribution to the force from the uniform background vanishes by
728: symmetry and thus, with this choice of the volume $V$, the force
729: coincides with the limit of the sum in Eq.\ref{force1}.
730:
731: While it is possible to show that the presence of an analogous
732: background with negative mass density comes naturally when the motion
733: of a particle is described in comoving coordinates in an expanding
734: universe, in pure Newtonian gravity such a background does not exist
735: and has to be introduced artificially to regularize the problem (Jean's
736: swindle). The negative mass density background is equivalent to the
737: condition that the force is summed symmetrically (see also
738: \cite{force,joyce}). Note that
739: this modification does not necessarily make the gravitational force
740: well defined in general: whether it is well defined depends on the
741: nature of the correlations between fluctuations in the density field
742: on large scales (see discussion in \cite{force}) .
743:
744:
745: Let us now consider the evolution from an initially cold (i.e. zero
746: velocity) Poisson distribution. Given that the PDF of the
747: gravitational force is well approximated by the NN one for large
748: fields, we can make a simple test: run a gravitational N-body
749: simulation starting from a cold Poisson distribution (nominally
750: infinite), with only this component of the force and then compare it
751: with the situation occurring in the case the full gravity force is
752: considered \cite{bsl04}. The approximate NN simulation is able to
753: reproduce the formation of the first structures, made by clusters of
754: two particles, up to the typical time scale $\tau_{NN}$ for two
755: particles initially placed at a distance $\ell$, the initial average
756: distance between NN, to collapse on each other. This is (again) of the
757: order of
758: \be
759: \label{nntime}
760: \tau_{NN} \sim \sqrt{G m/\ell^3}^{-1} \sim \sqrt{G \rho_0}^{-1} \;,
761: \ee
762: where $\rho_0 = m n_0$ and $n_0 \sim \ell^{-3}$ is the number
763: density. In particular it is very interesting to note that the form of
764: the two-point correlation function in the non-linear regime, which as
765: already mentioned will be preserved during time evolution because of
766: its ``self-similar'' behavior, forms during this NN phase. This is not
767: all however as the NN truncation will loose an important aspect of the
768: gravitational evolution, that on large scales, i.e. $r \gg \ell$,
769: small-amplitude density fluctuations are growing. This is an effect of
770: the long-range interactions, of weak amplitude, characterizing the
771: gravitational dynamics. Let us see briefly how can one model this
772: phenomenon.
773:
774:
775:
776:
777:
778: By considering the effect of the uniform negative background,
779: the Poisson equation becomes
780: \be
781: \nabla \vec{g} = - 4\pi G m (n(\vec{x}) -n_0)
782: \ee
783: where $\vec{g}$ is the gravitational field. By considering the
784: Vlasov-Poisson system of equations, after certain approximations, one
785: may derive the fluid equations describing the evolution of a
786: self-gravitating fluid (see e.g. \cite{peebles}). By performing a
787: perturbation analysis of these equations, for the case of
788: pressure-less matter around $\rho_0=mn_0$ and $\vec{\upsilon}=\vec{0}$
789: one finds at first order that the evolution of the density contrast
790: \[
791: \delta(\vec{x},t) = \frac{n(\vec{x,t}) -n_0}{n_0}
792: \]
793: is described, for the case in which the initial velocity is set equal
794: to zero, by
795: \be
796: \label{linear}
797: \delta(\vec{x},t) = \delta(\vec{x},0) \cosh\left( \sqrt{4 \pi G \rho_0} t
798: \right) =
799: \delta(\vec{x},0) \cosh\left(t/\tau \right)
800: \ee
801: where $\tau$ is the unique characteristic time scale of the system
802: which we have already mentioned in Eq.\ref{tau}. and which is of the
803: order of the NN collapse time scale given by Eq.\ref{nntime}, i.e. the
804: fluid dynamics and the NN dynamics are characterized by the same time
805: scale. This is the only time scale in the systems which will
806: characterize the evolution of all scales, from $\ell$ to the largest
807: scales in the system.
808:
809:
810: Eq.\ref{linear} describes one of the most basic results (see
811: e.g. \cite{peebles}) about self-gravitating systems, treated in a
812: fluid limit: that the amplitude of small fluctuations grows
813: monotonically in time, in a way which is independent of the
814: scale. This linearized treatment breaks down at any given scale when
815: the relative fluctuation $\delta$ at the same scale becomes of order
816: unity, signaling the onset of the ``non-linear'' phase of
817: gravitational collapse of the mass in regions of the corresponding
818: size. When the non-linear phase will involve many particles, the
819: objects formed will have properties similar to the core-halo structure
820: formed in the collapse of the finite system discussed above. These are
821: quasi stationary states which naturally emerge from the dynamical
822: evolution of an infinite self-gravitating system of particles starting
823: from quasi uniform initial conditions. These are the primary building
824: blocks in terms of which the non-linear structures observed in
825: cosmological simulations are described. The physical mechanism
826: driving the formation of these quasi-equilibrium states is similar to
827: the violent relaxation process described above, which in this case
828: occurs in more complicated situations (presence of sub-structures,
829: etc.) and environments (tidal effects from the boundary conditions,
830: etc.). Theoretically little is known about the dynamics
831: of these objects.
832:
833:
834: As discussed in more detail by \cite{joyce} (and see also
835: \cite{gravclusl,earlytimes}) the interplay between the small and
836: large scales, discrete and fluid dynamics, is an important element
837: for the comprehension of the formation of non-linear clustering in an
838: infinite system. Large scale fluctuations, i.e. linear theory,
839: determine the amplitude of correlations: in this way one expects
840: independence on discreteness for the relevant quantities as in the
841: fluid limit the discrete scale is sent to zero. On the other hand the
842: form of the self-similar two-point correlation function is established in the
843: transient phase dominated by two-body interactions, with a slightly
844: different time-dependence than fluid linear theory because of the
845: effect of discrete fluctuations \cite{plt1,plt2}.
846:
847: The considerations of these elements \cite{cg} (see also the
848: contribution by \cite{joyce} in this volume) lead to the formulation of a
849: simple model to explain the origin of ``self-similarity'' in the
850: gravitational clustering of an infinite distribution, where timescales
851: are dictated by the fluid limit but its non-linear dynamics are
852: intrinsically discrete. Given that discrete fluctuations characterize
853: any particle distribution, it is possible to understand why the
854: two-point correlation function is independent (or very weakly
855: dependent) on initial conditions.
856:
857:
858: It has been found \cite{univ} that non-linear clustering which results
859: in very different simulations is essentially the same, with a
860: well-defined simple power-law behavior in the two-point
861: correlations. This in fact a common feature of different sets of
862: gravitational N-body simulations with different initial particle
863: configurations, all describing the dynamics of discrete particles with
864: a small initial velocity dispersion (with or without the space
865: expansion of cosmological models). This apparently universal behavior
866: can be understood by the domination of the small scale contribution to
867: the gravitational force, coming initially from NN particles. In this
868: perspective the nature of clustering in the non-linear regime has
869: little to do with the initial fluctuations, or with the space
870: background being in expansion. Rather it is associated with what is
871: common to all these simulations: their evolution in the non-linear
872: regime is dominated by fluctuations on small scales, which are similar
873: in all cases at the time this clustering develops.
874:
875:
876: On the other hand is worth noticing that the usual theoretical
877: modeling (see e.g. \cite{peebles}) considers only a mean-field
878: approximation (Vlasov limit) and finds non universal behavior of the
879: two-point correlation function. In fact, in the cosmological
880: literature the idea is widely dispersed that the exponents in
881: non-linear clustering are related to that of the initial correlations
882: of the small fluctuations in the matter fluid, and even that the
883: non-linear two-point correlation can be considered an analytic
884: function of the initial two-point correlations
885: \cite{pd}. The models used to explain the behavior in the non-linear
886: regime usually involve both the expansion of the universe, and a
887: description of the clustering in terms of the evolution of a
888: continuous fluid. In this perspective the discrete nature of a
889: particle distribution is neglected. On the other hand discussion of
890: the small-dynamics at early time presented above already points toward
891: the fact that a complete characterization of the formation of
892: non-linear structures it is then required a careful study of the role
893: of discrete effects in the linear and non-linear regime (see also
894: \cite{joyce}).
895:
896:
897:
898:
899: \subsection{The infinite expanding case}
900:
901: Let us now turn to the more complex case of cosmological structure
902: formation. In this case one wants to understand gravitational
903: clustering in an expanding universe. One models this evolution
904: starting from the Friedmann solution of Einstein's field equations,
905: where the main assumption is that the matter density field is
906: considered as a perfectly uniform fluid at all scales
907: \cite{saslaw2,peebles}. It is then generally assumed that there is a
908: single length scale characterizing the correlation properties between
909: density fluctuations, which is of order $100$ Mpc\footnote{We refer
910: the interested reader to
911: \cite{book,glass,slv07} for a discussion about the correlation
912: properties of matter density fields in standard cosmological models.},
913: i.e. much smaller than radius of curvature of the universe (of order
914: $10^3$ Mpc) and that particles have velocities much smaller than the
915: velocity of light. In this situation one can treat the problem of
916: structure formation by assuming the approximation provided with
917: Newtonian mechanics in an expanding background (see Fig.\ref{fig6}).
918:
919: \begin{figure}
920: \includegraphics[height=.3\textheight]{fig6.ps}
921: \caption{Evolution of the fluctuations and formation of structures
922: in a simulation (with periodic boundary conditions, representing the
923: infinite system case) started from a correlated and cold initial
924: conditions. Even in this case structures form firstly on small scales
925: and then propagate to larger and larger scales. The initial conditions
926: are generated by applying a correlated displacement field to a
927: pre-initial perfect cubic lattice configuration (see
928: \cite{bsl02,jbruno}). One may note the appearance of ``filaments'', linking
929: clusters of particles of different sizes, marking the trace of
930: long-range correlations in the initial particle distribution. Indeed,
931: models of primordial density fields in the early universe predict the
932: presence of long-range weak amplitude correlations on small scales; on
933: large scales (i.e., $r>100$ Mpc) instead the matter distribution
934: should have super-homogeneous (or hyper-uniform) features (see
935: discussion in, e.g.,
936: \cite{book,glass,slv07}).}
937: \label{fig6}
938: \end{figure}
939:
940:
941: The equation of motion of particle in an expanding background is then
942: \be
943: \ddot{\vec x}_i + 2 H(t) \vec{x}_i = -\frac{1}{a^3} \sum_{i \ne j} Gm_j
944: \frac{\vec{x}_i -\vec{x}_j} {|\vec{x}_i -\vec{x}_j|^3}
945: \ee
946: where the dots dots denote derivatives with respect to time,
947: $\vec{x}_i$ is the comoving position of the $i^{th}$ particle, of mass
948: $m_i$, related to the physical coordinate by $\vec{r}_i =
949: a(t)\vec{x}_i$, and where $a(t)$ is the scale factor of the background
950: cosmology with Hubble constant $H(t) = \dot{a}/a$. The static case
951: can be obtained when $a(t)=1$ and thus $H(t) =0$. The expansion rate
952: $a(t)$ is determined by the type of energy component which dominates
953: the universe on large scales. For example one may consider a
954: matter-dominated universe or the case where expansion is dominated by
955: a cosmological constant (dark energy). In current cosmological models
956: the $70\%$ of the energy is made of dark energy, $25\%$ of dark matter
957: and $5 \%$ of ordinary baryonic matter.
958:
959:
960: In this situation one can develop a perturbation treatment of the
961: self-gravitating fluid equations in an expanding universe. The results
962: are similar with the static case discussed above and in a
963: matter-dominated universe linear theory describes a growing and a
964: decaying mode, both of them power laws in time rather than exponential
965: as in Eq.\ref{linear}. Thus in the linear regime at least, it is
966: possible to map the evolution in a static and in an expanding universe
967: by taking into account the different density fluctuations growth time
968: rates.
969:
970: In the non-linear regime, when cold initial conditions are considered,
971: one observes the same general characteristics found in the static
972: background case: a bottom-up aggregation process leading to a
973: ``self-similar'' evolution of the correlation function (in the sense
974: of Fig.\ref{fig4})\cite{efts88}. The time-rate of growth is power-law
975: instead of exponential also in the non-linear regime.
976:
977:
978: The simplest approach in modeling the large scale evolution into
979: non-linear matter structure is represented by the collapse of a single
980: over-dense region into a self-gravitating halo via the same spherical
981: collapse model used to study the collapse of a finite system (see e.g.
982: \cite{peacock,coles}). It is interesting to note that the radius of an
983: over-dense sphere behaves in the same way as the expansion factor for a
984: closed universe and therefore one is able to model the growth of a
985: spherically symmetric density perturbation using the same equations as
986: classical cosmology, i.e. the Friedmann equations.
987:
988: The collapse into self-gravitating virialized objects is then a common
989: feature to finite and infinite systems. In the context of cosmological
990: N-body simulations the core-halo structures are simply called halos,
991: by which it is intended the virialized part of these structures. Halos
992: are ubiquitous in N-body simulations and they show interesting
993: universal properties (density profiles, velocity distributions),
994: although different from the those of the finite system described
995: above. Even in this situation one would like to develop an analytic
996: treatment to understand the formation of these objects and their
997: properties. This is still lacking despite the fact that there have
998: been several attempts for an analytical derivation of the properties
999: of these halos, and in particular of their density and velocity
1000: profiles (see \cite{morikawa} in this volume). Standard approaches
1001: are usually based on non-collisional approximations (see
1002: \cite{hansen,white} for a discussion of the subject and for a list of
1003: references).
1004:
1005:
1006:
1007:
1008:
1009: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
1010:
1011: \section{Conclusions}
1012: \label{conclusions}
1013:
1014: The dynamics of infinite self-gravitating systems is a fascinating
1015: theoretical problem of out of equilibrium statistical mechanics,
1016: directly relevant both in the context of cosmology/astrophysics and,
1017: more generally, in the physics of systems with long-range
1018: interactions. We discussed some of the many problems encountered in
1019: the study of the gravitational clustering in both finite and infinite
1020: systems. We would like to stress two important open problems. The
1021: first concerns the extent to which such numerical simulations of a
1022: finite number of particles, reproduce the mean-field/Vlasov limit
1023: which is usually used to describe the evolution from a theoretical
1024: point of view. That is, the theoretical question that arises is about
1025: the validity of this collisionless limit. Another major question is
1026: that of the understanding of halo structure observed in simulations of
1027: infinite systems: while these show strongly universal characteristics,
1028: their dynamical origin is not yet understood
1029: from a theoretical point of view.
1030:
1031:
1032:
1033: \begin{theacknowledgments}
1034: I wish to thank Bruno Marcos, Michael Joyce and Andrea Gabrielli for
1035: useful comments and fruitful collaborations on the subject. I also
1036: thank Bill Saslaw for the very many discussions over the years we
1037: had together on this topic.
1038: \end{theacknowledgments}
1039:
1040:
1041: \begin{thebibliography}{9}
1042:
1043: \bibitem{intro} A. Campa, A. Giansanti, G. Morigi and F.
1044: Sylos Labini
1045: {\it in this volume}
1046:
1047:
1048: \bibitem{campa} A. Campa, {\it in this volume}
1049:
1050:
1051:
1052: \bibitem{chavanis} P.H. Chavanis, {\it in this volume}
1053:
1054: \bibitem{lyndebell} D. Lynden-Bell, Mon. Not. R. Astron. Soc.,
1055: {\bf 136}, 101, (1967).
1056:
1057: \bibitem{pad_physrep}
1058: T. Padmanabhan, Physics Reports, {\bf 188}, 285 (1990)
1059:
1060:
1061: \bibitem{chandra} S. Chandrasekhar, {\it ``Principles of Stellar Dynamics''},
1062: New York: Dover, 1960
1063:
1064: \bibitem{chandra_revmodphy} S. Chandrasekhar,
1065: Rev. Mod. Phys.,
1066: {\bf 15}, 1, (1943)
1067:
1068: \bibitem{pad_book} T. Padmanabhan,
1069: {\it ``Structure formation in the universe''},
1070: (Cambridge University Press, Cambridge, 1993)
1071:
1072: \bibitem{pad_dtslri}T. Padmanabhan in
1073: {\it ``Dynamics and Thermodynamics of Systems with Long-Range
1074: Interaction''}, Dauxois, S. Ruffo, E. Arimondo, M. Wilkens, eds., (Springer, Berlin 2002)
1075:
1076: \bibitem{saslaw1} W. C. Saslaw, {\it ``Gravitational Physics of Stellar and
1077: Galactic Systems''}, (Cambridge, UK: Cambridge Univ. Press,
1078: 1985)
1079:
1080: \bibitem{saslaw2} W. C. Saslaw,
1081: {\it ``The Distribution of the Galaxies: Gravitational Clustering in
1082: Cosmology''} (Cambridge, UK: Cambridge Univ. Press, 2000).
1083:
1084:
1085: \bibitem{binney}
1086: J. Binney and S. Tremaine, {\it ``Galactic Dynamics''} (Princeton
1087: University Press, 1994).
1088:
1089:
1090: \bibitem{peebles}
1091: P. J. E., Peebles, {\it The Large-Scale Structure of the Universe}
1092: (Princeton University Press, Princeton New Jersey, 1980)
1093:
1094:
1095: \bibitem{thierry} T. Baertschiger {\it Ph.D. Thesis}
1096: (University of Geneve, Geneva, Switzerland, 2004)
1097:
1098:
1099: \bibitem{peacock}J.A. Peacock,
1100: {\it ``Cosmological physics''}
1101: (Cambridge University Press, Cambridge, 1999)
1102:
1103:
1104: \bibitem{morikawa} M. Morikawa, {\it in this volume}
1105:
1106: \bibitem{joyce} M. Joyce, {\it in this volume}
1107:
1108: \bibitem{saslaw} W. C. Saslaw, {\it in this volume}
1109:
1110: \bibitem{binneyknebe}J. Binney and A.Knebe,
1111: Mon. Not. R. Astron. Soc., {\bf 333}, 378 (2002)
1112:
1113:
1114: \bibitem{diemond}
1115: J. Diemand, B. Moore, J. Stadel, and S. Kazantzidis,
1116: Mon. Not. R. Astron. Soc., {\bf 348}, 977 (2004)
1117:
1118: \bibitem{power02}
1119: C. Power et~al.
1120: Mon. Not. R. Astron. Soc.
1121: {\bf 338}, 14 (2003)
1122:
1123:
1124: \bibitem{violrelax}
1125: A. Arad and P.H. Johansson Mon. Not. R. Astron. Soc.
1126: {\bf 362}, 252 (2005)
1127:
1128: \bibitem{white} J.~F. Navarro,
1129: C.~S. Frenk,
1130: S.~D. White
1131: Astrophys. J., {\bf 490},
1132: 493 (1997)
1133:
1134:
1135: \bibitem{hansen} S. H. Hansen, B. Moore,
1136: New Astron. {\bf 11} 333 (2006)
1137:
1138:
1139: \bibitem{millenium}
1140: V. Springel, et al., Nature, {\bf 435}, 629 (2005)
1141:
1142: \bibitem{efts88}
1143: G. Efstathiou,
1144: C.~S. Frenk,
1145: S.~D.~M. White,
1146: and M. Davis,
1147: Mon. Not. R. Astr. Soc.,
1148: {\bf 235},
1149: 715 (1988).
1150:
1151: \bibitem{smith2003}
1152: R.~E. Smith et al.,
1153: Mon. Not. R. Astron. Soc.,
1154: {\bf 341}, 1311 (2003)
1155:
1156:
1157: \bibitem{book}
1158: A. Gabrielli, F. Sylos Labini, M. Joyce , L. Pietronero, {\it
1159: `` Statistical physics for cosmic structures''}, (Springer Verlag,
1160: Berlin, 2004)
1161:
1162:
1163:
1164: \bibitem{force} A. Gabrielli, M. Joyce, B. Marcos, F. Sylos Labini,
1165: T. Baertschiger,
1166: Phys. Rev. {\bf E74}, 021110 (2006)
1167:
1168: \bibitem{masucci} A. Gabrielli, P. A. Masucci \& F. Sylos Labini,
1169: Phys. Rev. {\bf E69}, 031110 (2004)
1170:
1171:
1172:
1173: \bibitem{pellegrini} A.
1174: Gabrielli, F. Sylos Labini, \& S. Pellegrini, Europhys.Lett., {\bf
1175: 46}, 127 (1999)
1176:
1177: \bibitem{bsl04} T. Baertschiger \& F. Sylos Labini,
1178: Phys. Rev., {\bf D69}, 123001-1 (2004)
1179:
1180: \bibitem{gravclusl} T. Baertschiger, A. Gabrielli, M. Joyce, F. Sylos
1181: Labini, Phys. Rev.{\bf E 75}, 059905 (2007)
1182:
1183: \bibitem{earlytimes}T. Baertschiger, M. Joyce, F. Sylos Labini and B. Marcos,
1184: {\tt arXiv:0711.2219}
1185:
1186: \bibitem{plt1} M. Joyce, B. Marcos, A. Gabrielli, T. Baertschiger,
1187: F. Sylos Labini, Phys. Rev. Lett., {\bf 95}, 011304 (2005)
1188:
1189: \bibitem{plt2} B. Marcos, T. Baertschiger, M. Joyce, A. Gabrielli,
1190: F. Sylos Labini, Phys. Rev., {\bf D73}, 103507 (2006)
1191:
1192:
1193: \bibitem{cg}T. Baertschiger, A. Gabrielli, M. Joyce, B. Marcos,
1194: F. Sylos Labini, Phys. Rev., {\bf E76}, 011116 (2007)
1195:
1196: \bibitem{univ} F. Sylos Labini, T. Baertschiger \& M. Joyce,
1197: Europhys.Lett., {\bf 66}, 171 (2004)
1198:
1199: \bibitem{glass} A. Gabrielli, M. Joyce,
1200: and F. Sylos Labini, Phys. Rev., {\bf D65}, 083523 (2002)
1201:
1202: \bibitem{slv07} F. Sylos Labini and N. L. Vasilyev
1203: Astron. Astrophys., in the press (2007) {\tt arXiv:0710.0224}
1204:
1205:
1206: \bibitem{bsl02} T. Baertschiger and F. Sylos Labini,
1207: Europhys.Lett., {\bf 57}, 322 (2002)
1208:
1209:
1210: \bibitem{jbruno}
1211: M. Joyce, D. Levesque, B. Marcos,
1212: Phys.Rev., {\bf D72}, 103509 (2005)
1213:
1214: \bibitem{pd}
1215: J. Peacock and S. Dodds, Mon. Not. R. Astron. Soc., {\bf 280},
1216: L19 (1996)
1217:
1218: \bibitem{coles} V. Sahni and
1219: P. Coles,
1220: Physics Reports, {\bf 262},
1221: 1 (1995).
1222:
1223:
1224:
1225:
1226:
1227: \end{thebibliography}
1228:
1229: \end{document}
1230:
1231:
1232: