1: \documentclass[prd,aps,floats,epsfig,eqsecnum,nofootinbib]{revtex4}
2: \usepackage{amsmath,amssymb,verbatim,epsfig,psfig,graphicx,rotating}
3: \newcommand{\be}{\begin{equation}}
4: \newcommand{\ee}{\end{equation}}
5: \newcommand{\bea}{\begin{eqnarray}}
6: \newcommand{\eea}{\end{eqnarray}}
7: \begin{document}
8: \title{\bf
9: THE SELF-GRAVITATING GAS IN THE PRESENCE OF DARK ENERGY: MONTE-CARLO
10: SIMULATIONS AND STABILITY ANALYSIS}
11: \author{ {\bf H. J. de Vega$^{(a,b)}$, J. A. Siebert$^{(a)}$\\}}
12: \affiliation{$^{(a)}$Laboratoire de Physique Th\'eorique et Hautes Energies, \\
13: Universit\'e Paris VI et Paris VII, Laboratoire Associ\'e au CNRS UMR 7589,
14: Tour 24, 5\`eme \'etage, \\ 4, Place Jussieu
15: 75252 Paris cedex 05, France.}
16: \affiliation{$^{(b)}$Observatoire de Paris, LERMA.
17: Laboratoire Associ\'e au CNRS UMR 8112.
18: \\61, Avenue de l'Observatoire, 75014 Paris, France.}
19: \begin{abstract}
20: The self-gravitating gas in the presence of a positive cosmological constant
21: $ \Lambda $ is studied in thermal equilibrium by Monte Carlo simulations and by
22: the mean field approach. We find excellent agreement between both approaches already
23: for $ N = 1000 $ particles on a volume $V$ [The mean field is exact in the infinite
24: $N$ limit]. The domain of stability of the gas is found to increase when the
25: cosmological constant increases. The particle density is shown to be an increasing
26: (decreasing) function of the distance when the dark energy dominates over
27: self-gravity (and vice-versa).
28: We confirm the validity of the thermodynamic limit: $ N, V \to \infty $
29: with $ N/V^{\frac13} $ and $\Lambda \; V^{\frac23} $ fixed.
30: In such dilute limit extensive thermodynamic quantities like energy, free energy,
31: entropy turn to be proportional to $N$.
32: We find that the gas is stable till the isothermal compressibility diverges.
33: Beyond this point the gas becomes a extremely dense object whose properties
34: are studied by Monte Carlo.
35: \end{abstract}
36: \date{\today}
37: \maketitle
38: \tableofcontents
39:
40: \section{Introduction}
41: The self-gravitating gas in thermal equilibrium has been thoroughly
42: studied since many years\cite{gasn,gas2,Emdsg,Chandrsg,Ebertsg,Bonnorsg,
43: Antonovsg,Lyndsg,Jeans,BT}. As a
44: consequence of the long range attractive Newton force, the
45: selfgravitating gas admits a consistent thermodynamic limit
46: $ N, V \to \infty $ with $ \frac{N}{V^{\frac13}} $ fixed. In this limit,
47: extensive thermodynamic quantities like energy, free energy, entropy
48: are proportional to $N$\cite{gasn,gas2}.
49:
50: In ref.\cite{gaslambda} we investigated how a cosmological constant
51: affects the properties of the non-relativistic self-gravitating gas
52: in thermal equilibrium by mean field methods. The mean field approximation
53: becomes exact in the limit when the number of particles becomes infinity.
54:
55: In the present paper we study the stability properties of the self-gravitating
56: gas in the presence of a cosmological constant by mean field and Monte-Carlo
57: methods.
58: The use of Monte-Carlo simulations is particularly useful since they are
59: like real experiments and allow to unambiguously determine the stability
60: or instability of the system.
61:
62: We compute both by Monte Carlo and mean field methods several physical
63: quantities: the equation of state, the particle density and the average
64: particle distance for different values of the dark energy. An excellent
65: agreement between both approaches is found except very close to the
66: collapse transition [see figs. 4-6]. The difference between both approaches
67: turns to be, as expected, of the order $ \frac1{N} $ where the number of
68: particles $N$ we choose in the Monte Carlo simulations was
69: $ N = 1000 - 2000 $. The slightly
70: larger split between Monte Carlo and mean field near the collapse is due to
71: the fact that there the corrections
72: to the mean field become singular \cite{gasn,gas2}.
73:
74: We find that the onset of instability in the canonical ensemble
75: coincides with the point where the isothermal compressibility diverges.
76: At this point we find that the dimensionless parameter $ \zeta =
77: \frac{G^3 \; m^6 N^2 \; P}{T^4} $ is maximal. ($ \zeta $ was introduced in
78: ref. \cite{Lyndsg} for $ \Lambda = 0 $). We find that the
79: domain of stability of the gas increases for increasing cosmological
80: constant. The dark energy has an anti-gravity effect that disfavours the
81: collapse pushing apart the particles.
82:
83: In absence of cosmological constant the particle density $\rho(r)$
84: is a {\bf decreasing}
85: function of the distance in the self-gravitating gas (see for example
86: \cite{gasn}). In the presence of a positive $ \Lambda $ we find that
87: $\rho(r)$ {\bf decreases} with $r$ when the self-gravity dominates
88: over the dark energy. This happens for $ X \equiv \frac{2 \Lambda V}{m N} < 1$.
89: For $X>1$ the cosmological constant dominates over the self-gravity and
90: the the particle density turns to {\bf increase} with the distance.
91: This is a consequence of the repulsive character of the dark energy.
92:
93: \medskip
94:
95: The Monte-Carlo study of the self-gravitating gas shows the validity of
96: the dilute thermodynamic limit introduced in refs.\cite{gasn,gaslambda}.
97: Namely, for $ N, V \to \infty $ with $ \frac{N}{V^{\frac13}} $ and
98: $\Lambda \; V^{\frac23} $ fixed, a consistent thermodynamic limit is reached
99: where extensive thermodynamic quantities like energy, free energy, entropy
100: are shown to be proportional to $N$.
101: Furthermore, the Monte-Carlo simulations allow us to study
102: the condensed phase which turns to be an extremely dense body.
103: The phase transition to collapse is found to be of zeroth order.
104:
105: \medskip
106:
107: The outline of the paper is as follows. Section II presents non-relativistic
108: selfgravitating particles in the presence of the cosmological constant,
109: their statistical mechanical treatment and the mean field approach.
110: In section III we analyze the stability of the selfgravitating gas
111: for zero and nonzero cosmological constant $\Lambda$
112: while in section IV we give the physical
113: picture of the gas and its particle density as a function of $\Lambda$.
114: Section V contains the results of our Monte-Carlo simulations on the phase diagram,
115: the collapse transition, the particle density and the properties of the collapsed
116: phase.
117:
118: \section{The self-gravitating gas in the presence of the dark energy}
119:
120: \subsection{Non-relativistic selfgravitating particles
121: in the presence of the cosmological constant}
122:
123: We recall some results about the self-gravitating gas in the presence of the
124: cosmological constant \cite{gaslambda}. In the Newtonian limit the Einstein
125: equations of general relativity become\cite{pee}
126: \begin{equation} \label{poiL}
127: \nabla^2 V = 4\pi \, G \, \rho - 8\pi \, G \, \Lambda \; .
128: \end {equation}
129: \noindent where $V$ stands for the gravitational potential, $\rho$ for the
130: density of massive particles and $\Lambda$ for the cosmological constant.
131: Eq.(\ref{poiL}) determines the weak and static gravitational field
132: produced by non relativistic matter in the presence of the cosmological
133: constant. For zero cosmological constant we recover the usual Poisson equation,
134: as it must be.
135:
136: The Hamiltonian for such set of self-gravitating particles in the presence of
137: the cosmological constant can be written in the center of mass frame
138: as\cite{gaslambda}
139: \begin{equation} \label{hamil}
140: H = \sum_i \frac{{\vec p_i}^2}{2 \, m_i^2} -G \; \sum_{i<j}
141: \frac{m_i \; m_j}{|{\vec q}_i-{\vec q}_j|}
142: - \frac{4\pi \, G \, \Lambda}{3} \; \sum_i m_i \; {\vec q}_i^{\,2} \quad ,
143: \end {equation}
144: where $ m_i $ stands for the mass of the particle at the point $ {\vec
145: q}_i $ with momentum ${\vec p_i} $ and
146: $$
147: \sum_i m_i \; {\vec q}_i = 0 \; ,
148: $$
149: since we choose center of mass coordinates.
150:
151: The cosmological constant contribution to the potential energy
152: grows negative for increasing values of the particle distances ${\vec q_i}$
153: to the center of mass.
154: Therefore, the gravitational effect of the cosmological
155: constant on particles amount to push them outwards.
156: It can be then said that the cosmological constant has an
157: {\bf anti-gravitational} effect.
158:
159: \subsection{Statistical mechanics of the self-gravitating gas with $ \Lambda
160: \neq 0$}
161:
162: We present here the statistical mechanics
163: of the self-gravitating gas in the presence of the cosmological
164: constant in the canonical ensemble. For simplicity we shall consider $N$
165: particles with identical mass $m$.
166:
167: The Hamiltonian is given by eq.(\ref{hamil}) and therefore
168: the classical partition function of the gas is then,
169: $$
170: Z(T,N,V)=\frac{1}{N!} \int\ldots\int \prod_{l=1}^{N}
171: \frac{{\rm d}^3 {\vec p_{l}} \; {\rm d}^3 {\vec q_{l}}}{(2 \pi)^3} \;
172: \; e^{- \frac{H}{T} } \; \;.
173: $$
174: \noindent It is convenient to introduce the dimensionless coordinates
175: ${\vec r_{l}}=\frac{{\vec q_{l}}}{L}$.
176: The momenta integrals are computed straightforwardly. Hence, the
177: partition function factorizes as the partition function of
178: a perfect gas times a coordinate integral.
179: \begin{equation}\label{part}
180: Z=\frac{V^{N}}{N!} \left(\frac {m T}{2 \pi}\right)^{3 N/2}
181: \; \int\ldots\int \prod_{l=1}^{N}{\rm d}^3{\vec r_{l}} \; \;
182: e^{\eta \, u_{P} +\frac{2 \pi}{3} \, \xi \, u_{N}} \; ,
183: \end{equation}
184: where
185: \begin{equation} \label{upun}
186: u_{P} \equiv \frac {1}{N} \sum_{1 \leq i < j \leq N}
187: \frac {1}{|{\vec r_{i}}-{\vec r_{j}}|} \quad , \quad
188: u_{N} \equiv \sum_{i=1}^{N} r_{i}^2
189: \; .
190: \end {equation}
191: \noindent The dimensionless parameters which characterize the strength of
192: self-gravity and dark energy are respectively,
193: \begin{equation} \label{etaxi}
194: \eta \equiv \frac{G\, m^2 \, N}{T \, L} \quad \mbox{and} \quad \xi \equiv
195: 2 \, \Lambda\, G\, m \, \frac{L^2}{T} \; .
196: \end {equation}
197: \noindent $ \eta $ can be interpreted as the ratio of the self-gravitating
198: energy per particle to the kinetic energy per particle.
199:
200: We define the ratio of dark energy and self-gravity
201: \begin{equation} \label{x}
202: X \equiv \frac{\xi}{\eta}=\frac{2 \Lambda V}{m N} .
203: \end{equation}
204: It is the ratio between the dark energy contained inside the volume $V$
205: and the mass of ordinary matter.
206:
207: In the large $N$ limit the $3N$-uple integral in eq.(\ref{part}) can
208: be approximated by a functional integral over the density.
209: In the saddle point approximation we find that the density
210: \begin {equation} \label{phi}
211: \rho({\bf x})=e^{\Phi({\bf x})} \; .
212: \end{equation}
213: obeys to the differential equation \cite{gaslambda}
214: \begin {equation} \label{Poisson}
215: \nabla^2 \Phi({\bf x})+4 \pi \; \left( \; \eta \; e^{\Phi({\bf x})}
216: -\xi \right)=0 \;.
217: \end{equation}
218: The density has to be normalized as
219: \begin {equation} \label{normalisation}
220: \int {\rm d}^3 {\bf x}\;\rho({\bf x})=1 \; \; .
221: \end{equation}
222: In hydrostatic equilibrium we obtain the same equation \cite{gaslambda}.
223: Therefore, hydrostatics and mean field are equivalent in the $
224: N \to \infty $ limit.
225:
226: \noindent In the limiting case $X=0$ eq.(\ref{Poisson}) becomes the well known
227: Lane-Emden equation in the absence of cosmological
228: constant (see refs. \cite{Emdsg,Chandrsg,Ebertsg,Bonnorsg,Antonovsg,Lyndsg,Katzsg,
229: Padmanabhansg,Saslawsg}).
230:
231: \subsection{The isothermal sphere with $ \Lambda
232: \neq 0$}
233:
234: \begin{figure}[htbp]
235: \rotatebox{-90}{\epsfig{file=f.ps,width=14cm,height=14cm}}
236: \caption{The density at the boundary $f(X,\eta)$ versus $\eta$ for
237: $ X=0,0.3,1,1.5 $ by the mean field approach. The instability points
238: $ \eta_o(X) $ where the thermal compressibility diverges are pinpointed
239: on the plot by a $+$.
240: They are $\eta_o(X=0)=1.510 \ldots$, $\eta_o(X=0.3)=1.63 \ldots$,
241: $ \eta_o(X=1)=2.04 \ldots$ and $\eta_o(X=1.5)=2.55\ldots$.
242: The gas is stable from $\eta=0$ till $\eta=\eta_o(X)$ in the
243: canonical ensemble. }
244: \label{ps}
245: \end{figure}
246:
247: For spherically symmetric configurations the mean field equation
248: (\ref{Poisson})
249: becomes an ordinary non-linear differential equation.
250: \begin {equation} \label{Poissonrad}
251: \frac{{\rm d}^2 \Phi}{{\rm d}R^2}+\frac{2}{R} \frac {{\rm d} \Phi}{{\rm
252: d}R} +4 \pi \; \left(\eta \; e^{\Phi(R)}-\xi \right) = 0 \; .
253: \end{equation}
254: \noindent
255: The various thermodynamic quantities are expressed in terms
256: of their solutions.
257: We work in a unit sphere, therefore the radial variable runs
258: in the interval $0 \leq R \leq R_{max} $,
259: $ R_{max} \equiv \left(\frac{3}{4 \pi}\right)^{\frac{1}{3}} $.
260: The density of particles $\rho(R)$ has
261: to be normalized according to eq.(\ref {normalisation}).
262: Integrating eq.(\ref{Poissonrad}) from $R=0$ to $R=R_{max} $ yields,
263: \begin{equation} \label{emx}
264: \eta-\xi=-R_{max}^2 \; \Phi'(R_{max})
265: \end{equation}
266: \noindent Setting,
267: \begin{equation} \label{transf}
268: \Phi(R)=u(x)+\ln{\frac{\xi^{R}}{\eta^{R}}} \quad , \quad x=\sqrt
269: {3 \xi^{R}} \; \frac{R}{R_{max}} \; ,
270: \end{equation}
271: \noindent
272: $\xi^{R}=\frac{\xi}{R_{max}}$ and $\eta^{R}=\frac{\eta}{R_{max}}$,
273: the saddle-point equation (\ref{Poissonrad}) simplifies as,
274: \begin{equation} \label{reducedeq}
275: \frac{{\rm d}^2 u}{{\rm d}x^2}+\frac{2}{x} \frac {{\rm d} u}{{\rm
276: d}x}+e^{u(x)}-1=0 \; .
277: \end{equation}
278: \noindent
279: In order to have a regular solution at origin we have to impose
280: \begin{equation} \label{up0}
281: u'(0)=0 \; .
282: \end{equation}
283: \noindent
284: We find from eqs.(\ref{emx})-(\ref{transf}) for fixed values of $\eta$
285: and $\xi$,
286: \begin{equation} \label{emx2}
287: u'\left(\sqrt{3 \xi^{R}} \right)=-\frac{\eta^{R}-\xi^{R}}{\sqrt{3 \xi^{R}}}
288: \; .
289: \end{equation}
290: Eqs.(\ref{up0}) and (\ref{emx2}) provide the boundary conditions for
291: the nonlinear ordinary differential equation (\ref{reducedeq}). In
292: particular, they impose the dependence of $u_0 \equiv u(0)$ on
293: $\eta^{R}$ and $\xi^{R}$.
294:
295: \noindent We find from eq.(\ref{reducedeq}) for small $x$ the
296: expansion of $u(x)$ in powers of $x$ with the result,
297: \begin{equation} \label{dl}
298: u(x)=u_0 + (1-e^{u_0}) \; \frac{x^2}{6}+e^{u_0} (e^{u_0}-1) \;
299: \frac{x^4}{120} + {\cal O}(x^6)\; .
300: \end{equation}
301: \noindent where we imposed eq.(\ref{up0}). The value of $u_0$ follows
302: by imposing eq.(\ref{emx2}).
303:
304: \noindent Using eqs.(\ref{x}), (\ref{phi}) and (\ref {transf})
305: we can express the
306: density of particles in terms of the solution of eq.(\ref {reducedeq})
307: \begin{equation} \label{dens}
308: \rho(R)=X \; \exp\left[u\left(\sqrt
309: {3 X \eta^{R}} \; \frac{R}{R_{max}}\right)\right]\; .
310: \end{equation}
311: \noindent We choose to express the physical quantities in terms of
312: $\eta$ [eq.(\ref{etaxi})] and the ratio $X$ of dark energy to self-gravity
313: [eq.(\ref{x})].
314: The local pressure obeys the ideal gas equation of state in local form
315: $ P(R)=\frac{N \; T}{V} \; \rho(R) \; . $
316: Using eqs.(\ref{dens}) the pressure at the boundary is given by
317: \begin{equation} \label{f}
318: P(R_{max})=\frac{N \; T}{V} \; X \; \exp\left[u\left(\sqrt {3
319: X \eta^{R} }\right)\right]\; \equiv \frac{N \; T}{V} \; f(X,\eta).
320: \end{equation}
321: \noindent The contrast $ C(X,\eta) $
322: between the pressure at the center and at the boundary can be written as,
323: $$
324: C(X,\eta)\equiv\frac{P(0)}{P(R_{max})} \; =e^{u_0-u\left(\sqrt
325: {3 X \eta^{R}}\right)} \; .
326: $$
327: \noindent We plot in fig. \ref{ps} the density at the boundary
328: versus $\eta$ for different values of $X$.
329:
330: \section{Stability of the self-gravitating gas}
331:
332: \begin{figure}[htbp]
333: \rotatebox{-90}{\epsfig{file=etao.ps,width=14cm,height=14cm}}
334: \caption{The value of the parameter $\eta$ at the Jeans instability:
335: $\eta=\eta_o(X)$ versus $X$.}
336: \label{etao}
337: \end{figure}
338:
339: \subsection{Stability for $\Lambda=0$.}
340:
341: The self-gravitating
342: system can be in one of two phases: gaseous or highly condensed.
343: Mean field theory describes the gas phase and gives exactly the
344: physical quantities in the thermodynamic limit.
345: The equilibrium density $\rho({\vec x})$ solution of the mean field
346: equation (\ref{Poisson})
347: minimizes the Helmholtz free energy in the {\bf canonical} ensemble.
348: This density is the most probable distribution which becomes absolutely
349: certain in the infinite $N$ limit. All thermodynamics quantities follows
350: from this density $\rho({\vec x})$. This continuous density $\rho({\vec x})$
351: solution of the mean field equation (\ref{Poisson}) fails to describe the
352: condensed phase.
353:
354: The condensed phase have been found by Monte-Carlo
355: simulations for $\Lambda=0$ \cite{gasn}. In the microcanonical ensemble the
356: center collapses in a very dense core surrounded by a halo of particles. In
357: the canonical ensemble all the particles collapse in a very dense body
358: \cite{gasn}.
359:
360: The stability of the gaseous phase in the mean field
361: approach can be analyzed looking for the extrema of the dimensionless
362: parameter\cite{Lyndsg}
363: \begin{equation}
364: \label{zeta}
365: \zeta=\frac{G^3 \; m^6 N^2 \; P}{T^4} \; .
366: \end{equation}
367: when the temperature $T$, the pressure $P$ and the number of particles $N$ are
368: kept fixed in the canonical ensemble. The gas phase becomes instable
369: (for $\Lambda=0$) when $\zeta$ is maximal\cite{Lyndsg}.
370:
371: Using eqs.(\ref{etaxi}), (\ref{f}) and (\ref{zeta}) we find that
372: \begin{equation}\label{zetaet}
373: \zeta=\eta^3 \; f_0(\eta) \; .
374: \end{equation}
375: where $f_0(\eta) \equiv f(X=0,\eta)$ is the external density in absence of
376: dark energy.
377:
378: A better physical insight is obtained by looking to the the isothermal
379: compressibility
380: \begin{equation}
381: \label{defkt}
382: K_T=-\frac{1}{V} \; \left( \frac{\partial V}{\partial P}
383: \right)_{T} \; .
384: \end{equation}
385: \noindent Using eqs. (\ref{etaxi}) and (\ref{f}) we find that the
386: dimensionless isothermal compressibility
387: $\kappa_T \equiv \frac{NT}{V} \; K_T$ takes the form
388: \begin{equation} \label{kappat}
389: \kappa_T = \frac{1}{f_0(\eta)+\frac{\eta }{3} f_0^{'}(\eta)} \; .
390: \end{equation}
391: We see from eqs.(\ref{zetaet}) and (\ref{kappat}) that
392: $$
393: \kappa_T = \frac{3 \; \eta^2}{\frac{d\zeta}{d\eta}} \; .
394: $$
395: Hence, $\kappa_T$ {\bf diverges} at the extrema of $ \zeta $.
396:
397: The isothermal compressibility $\kappa_T$ is positive from $\eta=0$
398: till $\eta=\eta_o=1.510 \ldots$. At this point $\kappa_T$ as well
399: as the specific heat at constant pressure diverge and change
400: their signs\cite{Ebertsg,gasn}. Moreover, at this point
401: the speed of sound at the center of the sphere becomes imaginary\cite{gas2}.
402: Therefore, small density
403: fluctuations will grow exponentially in time instead of exhibiting oscillatory
404: propagation. Such a behaviour leads to the Jeans instability and collapse\cite{Jeans}.
405: Monte-Carlo simulations confirmed the presence of this instability at
406: $\eta=\eta_o=1.510 \ldots$ in the canonical ensemble for $\Lambda=0$
407: \cite{gasn}.
408:
409: \subsection{Jeans instability for $\Lambda \neq0$}
410:
411: We now compute the dimensionless isothermal compressibility for the
412: self-gravitating gas with $\Lambda \neq 0$.
413: Introducing the dimensionless parameter
414: \begin{equation} \label{alpha}
415: \alpha=\frac{2 \; G^3 \; m^5 \; N^2 \Lambda}{T^3} \; = \eta^3 \; X ,
416: \end{equation}
417: the dimensionless isothermal compressibility $ \kappa_T $ takes the form:
418: \begin{equation} \label{kappatlamb}
419: \kappa_T = \frac{1}{f(\eta,\alpha)+\frac{\eta }{3} \;
420: \left(\frac{\partial f}{\partial \eta} \right) (\eta,\alpha)} \; .
421: \end{equation}
422: \noindent The Jeans instability happens when the isothermal compressibility
423: diverges. This happens for $\eta=\eta_o(X)$ verifying:
424: $$
425: f(\eta_o,\alpha)+\frac{\eta_o }{3} \;
426: \left(\frac{\partial f}{\partial \eta} \right) (\eta_o,\alpha) = 0 \; .
427: $$
428: Thus, for a fixed $\alpha$, $\eta_o$ is the value of $\eta$ which
429: maximizes the quantity
430: \begin{equation} \label{zetax}
431: \zeta \equiv \eta^3 \; f(\eta,\alpha) =\frac{G^3 \; m^6 \; N^2 \; P}{T^4} \; .
432: \end{equation}
433: \noindent As for the self-gravitating gas with
434: $\Lambda=0, \; \zeta$ is maximal when the specific heat at constant
435: pressure diverges and changes its sign.
436: We plot $\eta_o$ versus $X$ in fig. \ref{etao} using eq.(\ref{alpha}).
437: Monte Carlo simulations confirm that the onset of the Jeans collapse happens
438: at $ \eta = \eta_o $.
439:
440: We can write $ \eta $ [eq.(\ref{etaxi})] as
441: $$
442: \eta = \frac{G\, m \, M}{T \, L}
443: $$
444: where $ M $ is the total mass of the gas. Therefore, the collapse
445: condition $ \eta \geq \eta_o(X) $ can be written as
446: $$
447: M \geq M_0 \; \eta_o(X) \quad {\mbox where} \quad M_0 \equiv
448: \frac{L \, T}{G \, m} \; .
449: $$
450: The gas collapses when its mass takes a larger value than the critical one $ M_J(X)
451: \equiv M_0 \; \eta_o(X) $. We can consider $ M_J(X) $ as the generalization
452: of the Jeans mass in the presence of the cosmological constant. Notice that
453: the presence of $ \Lambda $ {\bf increases } the Jeans mass with respect the
454: $ \Lambda = 0 $ case [see fig. \ref{etao}].
455:
456: \section{Physical Picture}
457:
458: \subsection{Behaviour of the gas phase with X}
459:
460: The effects of self-gravitation and dark energy go in opposite directions.
461: Self-gravitating
462: forces are attractive, while dark energy produces repulsion. If the ratio
463: of dark energy and self-gravity is
464: $X<1$ [defined by eq.(\ref{x})], self-gravitation dominates over dark energy.
465: If $X=1$ the effect of self-gravitation is exactly compensated by the
466: dark energy. If $X>1$ dark energy dominates over self-gravitation.
467: We illustrate these three cases plotting $f(X,\eta)$ versus $\eta$ for
468: $X=0.3,1,1.5$ in fig. \ref{ps}
469:
470: \begin{itemize}
471:
472: \item {\bf first case: $X<1$: self-gravity dominates}
473:
474: We illustrate this case choosing $X=0.3$. We see in fig. \ref{ps} that
475: $f(X,\eta)$ versus $\eta$ exhibits a form analogous to
476: $f(X=0,\eta)$ versus $\eta$ in absence of dark energy. $f(X,\eta)$ has two
477: Riemann sheets as a function of $\eta$.
478: In the first sheet $f(X,\eta)$ monotonically decreases with
479: $\eta$ for fixed $ X < 1 $.
480:
481: \item {\bf second case: $X=1$: exact compensation}
482:
483: The effect of self-gravitation is compensated by the effect of dark energy.
484: We see in fig. \ref{ps} that $f(X=1,\eta)=1$.
485: An exactly homogeneous sphere $\rho(R)=1$ is a solution (see below).
486: In such special case the self-gravitating gas behaves as a perfect gas
487: (with $PV=NT$ everywhere).
488:
489: \item {\bf third case: $X>1$: dark energy dominates}
490:
491: We illustrate this case choosing $X=1.5$. We see in fig. \ref{ps} that
492: $f(X=1.5,\eta)$ is an increasing function of $\eta$ for fixed $ X > 1 $.
493:
494: \end{itemize}
495:
496: \noindent The Jeans instability happens in the three cases. The value of
497: $ \eta = \eta_o(X)$ where the gas collapses increases with $X$
498: (fig. \ref{etao}). The gas collapses for $ X>0$ at a higher density
499: $\frac{N}{V^{\frac{1}{3}}}$ for a given temperature or for a lower temperature
500: for a given density $\frac{N}{V^{\frac{1}{3}}}$ than for $X=0$. That is, the
501: domain of stability of the gas increases for increasing $X$. The dark energy
502: has an anti-gravity effect that disfavours the collapse pushing the particles
503: toward the boundary of the sphere. Notice that the point of Jeans instability
504: is in the first Riemann sheet of the $\eta$-plane for $0 \leq X<1$.
505:
506: \subsection{The particle density $\rho(R)$}
507:
508: \begin{figure}[htbp]
509: \rotatebox{-90}{\epsfig{file=densksi.ps,width=14cm,height=14cm}}
510: \caption{The particle densities $\rho(R)$ vs. the radial coordinate $R$
511: for $X=0.3$ and $\eta = 1$ (decreasing), $X=1$ and $\eta = 1$ (homogeneous)
512: and finally $X=1.5$ and $\eta =1$ (increasing) computed by the mean field
513: approach}
514: \label{rho}
515: \end{figure}
516:
517: We study the density $\rho(R)$ as a function of the radial coordinate $R$. The
518: density $\rho(R)$ of the self-gravitating gas in absence of dark
519: energy ($X=0$) is always a {\bf decreasing} function of $R$. This follows
520: from the attractive character of gravitation (see for example ref.
521: \cite{gasn,gas2}).
522:
523: The $R$-dependence of the density
524: $\rho(R)$ of a self-gravitating gas in the presence of dark energy is more
525: involved because the dark energy opposes to the attraction by gravity.
526: The functions $\rho(R)$ and $u(x)$ have the same qualitative dependence in
527: $R$ since they are related by exponentiation [see eq. (\ref{dens})].
528:
529: The behaviour at the center of the sphere ($R=0$) is
530: governed by the sign of $u_0$ since $u'(0)=0$ [eq. (\ref{up0})] and we find
531: from eq.(\ref{dl}) that $ u''(0)= \frac{1}{3} \!
532: \left(1-e^{u_0} \right)$. Hence,
533: $$
534: \mbox{sign}[ u''(0) ] = -\mbox{sign}[ u_0 ] \; .
535: $$
536: Therefore, $\rho(R)$ decreases (increases) at $R=0$ if $u_0>0$ ($u_0<0$).
537:
538: The behaviour at the boundary of the sphere
539: ($R=R_{max}$) is governed by the sign of $\eta-\xi$ since according to
540: eq.(\ref{emx2})
541: $$
542: \mbox{sign}[ u'(R_{max}) ] = -\mbox{sign}[ \eta^{R}-\xi^{R} ] =
543: \mbox{sign}(X-1) \; .
544: $$
545: Therefore, $\rho(R)$ decreases (increases) at $R_{max}$ if $X<1$ ($X>1$).
546:
547: \medskip
548:
549: The particle density exhibits for the stable solutions (namely, for
550: $\kappa_T > 0$) one of the following behaviours illustrated in fig. \ref{rho}
551:
552: \begin{itemize}
553:
554: \item{\bf decreasing}: $X<1$ and $u_0>0$. The density $\rho(R)$ decreases
555: from the center of the sphere till the boundary. We plot in fig. \ref{rho}
556: the density $\rho(R)$ versus $R$ for $X=0.3$ and $\eta=1$. The self-gravitation
557: dominates over dark energy.
558:
559:
560: \item{\bf increasing}: $X>1$ and $u_0<0$. The density $\rho(R)$
561: increases from the center of the sphere till the boundary.
562: We plot in fig. \ref{rho}
563: the density $\rho(R)$ versus $R$ for $X=1.5$ and $\eta=1$. The dark energy
564: dominates over self-gravitation.
565:
566: \item{\bf homogeneous}: $X=1$ and $u_0=0$. The density is homogeneous:
567: $\rho(R)=1$. The effect of self-gravity of particles is exactly compensated by
568: the effect of dark energy. We plot in fig. \ref{rho}
569: the density $\rho(R)$ versus $R$ for $X=1$ and $\eta=1$.
570: \end{itemize}
571:
572: Eq.(\ref{Poisson}) for $ X=1 $ is similar to the equation that describes
573: in two space dimensions (multi)-vortices in the
574: Ginsburg-Landau or Higgs model in the limit between superconductivity
575: of type I and II \cite{vortex}. However, for the vortex case one has $
576: \xi = \eta < 0 $ since like charges repel each other in electrodynamics
577: while masses attract each other in gravity. This opposite nature of the
578: forces is responsible of the fact that eq.(\ref{Poisson}) for $ X=1 $ only
579: has trivial stable solutions while a host of non-trivial solutions
580: appear in the vortex case\cite{vortex}.
581:
582: \section{Monte Carlo calculations}
583:
584: \subsection{The Metropolis algorithm}
585:
586: The standard Metropolis algorithm \cite{MC} was first applied to
587: self-gravitating gas in absence of dark energy in ref.\cite{gasn}.
588: We apply this standard Metropolis to the self-gravitating gas
589: in presence of dark energy. We perform it in a volume $V$ in the
590: canonical ensemble at temperature $T$. We compute in this way the external
591: pressure, the energy, the average density, the average particle distance and
592: the average squared particle distance as function
593: of $\eta$ for a given value of the ratio of dark energy
594: and self-gravity $X \equiv \frac{\xi}{\eta}=\frac{2 \Lambda V}{m N}$.
595:
596: We implement the Metropolis algorithm in the following way. We start from a
597: random distribution of $N$ particles in the chosen volume. We update such
598: configuration choosing a particle at random and
599: changing at random its position. We then compare the energies of the former and
600: the new configurations.
601: We use the standard Metropolis test to choose between the new and the former
602: configurations. The energy of the configurations
603: are calculated performing the exact sums as in eq.(\ref{upun}). We use as
604: statistical weight for the Metropolis algorithm in the canonical ensemble,
605: $$
606: e^{\eta \, u_{P} +\frac{2 \pi}{3} \, \xi \, u_{N}} \; ,
607: $$
608: [see eqs.(\ref{part}) and (\ref{upun})].
609: The number of particles go up to $2000$.
610:
611: We introduce a short distance cutoff in the Newtons' potential
612: [see eq.(\ref{upun})]
613: $$
614: u_{P} \equiv \frac {1}{N} \sum_{1 \leq i < j \leq N}
615: \frac {1}{|{\vec r_{i}}-{\vec r_{j}}|_A} \; ,
616: $$
617: with
618: \begin{eqnarray}
619: |{\vec r_{i}}-{\vec r_{j}}|_A
620: &=&|{\vec r_{i}}-{\vec r_{j}}| \quad , \quad
621: \mbox{for} \quad |{\vec r_{i}}-{\vec r_{j}}| > A \nonumber\\
622: &=&A \quad , \quad \mbox{for} \quad |{\vec r_{i}}-{\vec r_{j}}|
623: < A \; , \nonumber
624: \end{eqnarray}
625: \noindent where $A \ll V^{\frac{1}{3}}$ is the short distance cut-off.
626: The presence of the short distance cut-off prevents the collapse
627: (here unphysical) of the self-gravitating gas (for more details see
628: \cite{gasn}).
629:
630: \subsection{Influence of the geometry}
631:
632: We apply the Metropolis algorithm to the self-gravitating gas in presence
633: of dark energy in a sphere and in a cube.
634:
635: \noindent For $X=0$ (absence of dark energy) the Monte Carlo results
636: in a sphere and in a cube with identical volume
637: give practically the same results \cite{gasn}.
638: The geometry does not influence the physics of the system.
639:
640: \noindent However, for $X>0$ (presence of dark energy) the Monte Carlo results
641: in a sphere and in a cube turn to be different. This effect can be traced
642: to the repulsive character of the dark energy that pushes the particles towards
643: the boundary. As a result, the physical quantities becomes sensible to the
644: geometry. We thus compared the mean field results with spherical symmetry
645: from the previous sections and ref.\cite{gaslambda} with Monte Carlo
646: calculations performed on a sphere.
647:
648: \subsection{Phase diagrams}
649:
650: \begin{figure}[htbp]
651: \rotatebox{-90}{\epsfig{file=fmc0p3.ps,width=14cm,height=14cm}}
652: \caption{The density at the boundary $f_X(\eta)=\frac{PV}{NT}$ versus $\eta$
653: for $X=0.3$ by Monte Carlo methods and mean field approach. The collapse value
654: of the gas phase is $\eta_T(X=0.3)=1.63\ldots$}
655: \label{fmc0p3}
656: \end{figure}
657:
658: \begin{figure}[htbp]
659: \rotatebox{-90}{\epsfig{file=r.ps,width=14cm,height=14cm}}
660: \caption{The average particle distance and the average squared particle
661: distance versus $\eta$ for
662: $X=0.3$ by Monte Carlo methods. They exhibit a
663: sharp decrease at the collapse value $\eta=\eta_T(X=0.3)=1.63\ldots$.}
664: \label{rmc}
665: \end{figure}
666:
667: \begin{figure}[htbp]
668: \rotatebox{-90}{\epsfig{file=fmc1p5.ps,width=14cm,height=14cm}}
669: \caption{The density at the boundary $f_X(\eta)=\frac{PV}{NT}$ versus $\eta$
670: for $X=1.5$ by Monte Carlo methods and mean field approach. The collapse value
671: of the gas phase is $\eta_T(X=1.5)=2.6\ldots$}
672: \label{fmc1p5}
673: \end{figure}
674:
675: We recall that our self-gravitating system exhibits three different
676: behaviours according the value of $X$,
677: the relative rate of dark energy and self-gravity.
678: In all three cases two different phases show up:
679: for $\eta < \eta_T(X)$ we have a non perfect gas and for $\eta > \eta_T(X)$
680: it is a very condensed system with negative pressure. The transition between the
681: two phases is very sharp.
682:
683: \begin{itemize}
684:
685: \item For $X<1$ the self-gravity dominates over dark energy.
686: We plot the external density $f(X,\eta)= \frac{PV}{NT}$ versus $\eta$
687: for $X=0.3$ in fig. \ref{fmc0p3}. In the gas phase
688: $\frac{PV}{NT}$ monotonically decreases with $\eta$ from $\eta=0$
689: till the collapse point $\eta=\eta_T(X)$.
690:
691: \noindent The average distance between particles $<|{\vec r}_i - {\vec r}_j|>$
692: and the average squared distance between particles
693: $\sqrt{<|{\vec r}_i - {\vec r}_j|^2>}$ monotonically decrease with
694: $\eta$ (fig. \ref{rmc}). When the gas collapses at $\eta_T$,
695: $<|{\vec r}_i - {\vec r}_j|>$ and $\sqrt{<|{\vec r}_i - {\vec r}_j|^2>}$
696: exhibit a sharp drop.
697:
698: \item For $X=1$ the effect of dark energy exactly compensates the effect of
699: self-gravity. We have $\frac{PV}{NT}=1$ in the gas phase.
700: The system behaves like a perfect gas from $\eta=0$
701: till the collapse point $\eta=\eta_T(X)$.
702:
703: \item For $X>1$ the dark energy dominates over self-gravity.
704: We plot in fig. \ref{fmc1p5} the external density
705: $f(X,\eta)=\frac{PV}{NT}$ versus $\eta$ for $X=1.5$.
706: In the gas phase $\frac{PV}{NT}$ monotonically increases with $\eta$
707: from $\eta=0$ till the collapse point $\eta=\eta_T(X)$.
708:
709: \end{itemize}
710:
711: We find that the Monte Carlo calculations in the sphere
712: accurately reproduce the mean field results for $\frac{PV}{NT}$ in the
713: gas phase with spherical symmetry.
714: We then consider a cube and a sphere of identical volume. The
715: Monte Carlo calculations turn to give different results in these two cases:
716: a) The external pressure is larger in the cube than in the sphere.
717: b) The gas phase is more stable in the cube. c) The collapse value $\eta_T(X)$
718: in the cube is larger than its value for the same $X$ in the sphere.
719:
720: \subsection{The collapse point}
721:
722: The phase transition happens in the Monte Carlo simulations
723: at $\eta=\eta_T(X)$. It must be compared with
724: the point of Jeans instability $\eta=\eta_o(X)$ according to mean field
725: [see III.C] where the isothermal compressibility diverges and changes its sign.
726: We see in that $\eta_T(X)$ is {\bf very close} of $\eta_o(X)$.
727: They are probably the same point. For $X=0.3$ we have $\eta_o(X)=1.63 \ldots$
728: and $\eta_T(X)=1.63 \ldots$.
729: For $X=1.5$ we have $\eta_o(X)=2.55 \ldots$
730: and $\eta_T(X)= 2.6\ldots$.
731: We conclude that the collapse of the system observed in the Monte Carlo
732: simulations is due to the Jeans instability when the isothermal compressibility
733: diverges and is well described by the mean field approach.
734:
735: For $\eta \gtrsim \eta_T(X)$ we see that the gaseous phase is metastable.
736: Before collapsing the system stay for a long Monte Carlo time in the
737: metastable gaseous phase in the Monte Carlo computations.
738: The Monte Carlo time for collapse increases with
739: increasing value of $X$. This is due to the repulsive effect of dark energy.
740: The gas phase is less unstable in the presence of dark energy.
741:
742: \subsection{Average distribution of particles}
743:
744: We illustrate the two behaviours of the self-gravitating gas in the presence
745: of the cosmological constant plotting
746: the average density as a function of two Cartesian coordinates, the third
747: coordinate being fixed. For simplicity we depict the densities in the cube.
748:
749: \noindent For $X>1$ we see that the density is larger on the boundary that
750: at the center of the cube (fig. \ref{densinc}). The dark energy dominates over
751: the self-gravity and pushes the particles toward the boundary of the cube.
752:
753: \noindent For $X<1$ we see that the density is larger at the center than on
754: the boundary of the cube (fig. \ref{densdec}).
755: The self-gravity dominates over dark energy and attracts the particles to the
756: center of the cube.
757:
758: We illustrate now the two phases (gaseous and condensed)
759: plotting the average particle distribution in a cube.
760: Fig. \ref{profgas} and fig. \ref{profcoll} depict the average particle
761: distribution
762: from Monte Carlo calculations with $1000$ particles for $X=0.3$ at both sides
763: of the collapse point. Fig. \ref{profgas} corresponds
764: to the gaseous phase and fig. \ref{profcoll} to the collapsed phase.
765:
766: We find that the Monte Carlo simulations (describing thermal equilibrium)
767: are much more efficient than the $N$-body simulations integrating Newton's
768: equations of motion. [Indeed, the integration of Newton's equations
769: provides much more information than thermal equilibrium investigations].
770: Actually, a few hundreds of particles are enough to get
771: quite accurate results in the Monte Carlo simulations (except near the collapse
772: points). Moreover, the Monte Carlo results turns to be in excellent
773: agreement with the mean field calculations up to corrections of the
774: order $ \frac1{N} $.
775:
776: \subsection{The condensed phase}
777:
778: In the condensed phase all the particles collapse in a very dense body
779: (see fig. \ref{profcoll}). The self-gravity contribution of the potential
780: energy dominates overwhelmingly the dark energy contribution in the condensed
781: phase. Using eqs. (\ref{hamil}), (\ref{etaxi}) and (\ref{x})
782: and the virial theorem \cite{gaslambda}
783: we obtain that the external pressure is expressed as
784: $$
785: f_X(\eta)=\frac{PV}{NT}=1-\frac{1}{3}
786: \left< \frac{1}{|{\vec r}_i-{\vec r}_j|} \right> \; \eta \; .
787: $$
788: The Monte Carlo results indicate that
789: $\left< \frac{1}{|{\vec r}_i-{\vec r}_j|} \right> \simeq 50$.
790: Thus in this condensed phase the external pressure can be
791: approximated by
792: $$
793: f_X(\eta)=\frac{PV}{NT}=1-K \eta \; .
794: $$
795: where $K \simeq 16$.
796:
797: Since $ f(\eta) $ has a jump at the transition, the Gibbs free energy
798: is discontinuous and we have a phase transition of the {\bf zeroth} order
799: as in the absence of cosmological constant \cite{gasn}.
800:
801: \section{Discussion and Conclusions}
802:
803: The behaviour of the self-gravitating gas is significantly influenced
804: by the cosmological constant (or not) depending on the value of the ratio
805: $X$ defined by eq.(\ref{x}),
806: $$
807: X = \frac{2 \, \Lambda \; V}{m \; N} = 2 \;
808: \left(\frac{\mbox{cosmological \; constant}}{\mbox{mass}}\right)_V = 2 \;
809: \frac{\Lambda}{\rho_V}
810: $$
811: where $\rho_V \equiv N \; m /V $ is the average mass density in the
812: volume $V$ and where the subscript $\left( \right)_V$ indicates mass and
813: cosmological constant inside the volume $V$.
814:
815: $X$ takes small values for astrophysical objects except for clusters
816: of galaxies where $ X{galaxy clusters} \sim 0.01 $ and especially for the universe
817: as a whole where $ X{universe} \sim 4 $.
818: The smallness of $X$ stems from the tiny value of the present cosmological constant
819: $\Lambda = 0.663 \; 10^{-29}$g/cm$^3$\cite{lam}. The particle mass density is
820: much larger than $\Lambda$ in all situations except for the universe as a
821: whole.
822:
823: However, the non-relativistic and equilibrium treatment does not
824: apply for the universe as a whole.
825: Certainly, the galaxy distribution can be considered as a non-relativistic
826: self-gravitating gas in the presence of the cosmological constant.
827: But the gas of galaxies is certainly not yet at thermal equilibrium today.
828: Perhaps, thermal equilibrium may be reached in some billions of years.
829: Incidentally, at thermal
830: equilibrium the particle density is regular at the center of the distribution
831: in the whole range of $ X $. Namely, there is no cusp at the center
832: in a situation of thermal equilibrium.
833:
834:
835: \begin{figure}[htbp]
836: \rotatebox{-90}{\epsfig{file=densx3eta1z5.ps,width=14cm,height=14cm}}
837: \caption{The local density of the gas $\rho(x,y,z)$ versus
838: the Cartesian coordinates $x$ and $y$ at $z=5$ in a cube of size $9$
839: for $X=3$ and $\eta=1$ from Monte Carlo simulations.
840: The density is larger on the boundary of the cube than at the center. }
841: \label{densinc}
842: \end{figure}
843:
844: \begin{figure}[htbp]
845: \rotatebox{-90}{\epsfig{file=densx0p3eta1z5.ps,width=14cm,height=14cm}}
846: \caption{The local density of the gas $\rho(x,y,z)$ versus
847: the Cartesian coordinates $x$ and $y$ at $z=5$ in a cube of size $9$
848: for $X=0.3$ and $\eta=1$ from Monte Carlo simulations.
849: The density is larger at the center of the cube than on the boundary.}
850: \label{densdec}
851: \end{figure}
852:
853: \begin{figure}[htbp]
854: \rotatebox{-90}{\epsfig{file=profx0p3eta1p6.ps,width=14cm,height=14cm}}
855: \caption{Average particle distribution in the gaseous phase from Monte
856: Carlo simulations in a cubic volume for $X=0.3$, $\eta=1.6$ and $N=2000$.
857: A $+$ denotes one particle.}
858: \label{profgas}
859: \end{figure}
860:
861: \begin{figure}[htbp]
862: \rotatebox{-90}{\epsfig{file=profx0p3eta1p8.ps,width=14cm,height=14cm}}
863: \caption{Average particle distribution in the collapsed phase from Monte
864: Carlo simulations in a cubic volume for $X=0.3$, $\eta=1.8$ and $N=2000$.
865: A $+$ denotes one particle.}
866: \label{profcoll}
867: \end{figure}
868:
869: \begin{thebibliography}{99}
870:
871: \bibitem{pee} P. J. E. Peebles, Principles of Physical Cosmology,
872: Princeton (1993).
873:
874: P. J. E. Peebles, B. Ratra, Revs. Mod. Phys. 75, 559 (2003).
875:
876: \bibitem{gasn} H. J. de Vega, N. S\'anchez, Phys. Lett. {\bf B490}, 180
877: (2000).
878:
879: H. J. de Vega, N. S\'anchez, Nucl. Phys. {\bf B 625}, 409 (2002).
880:
881: \bibitem{gas2} H. J. de Vega, N. S\'anchez, Nucl. Phys. {\bf B 625},
882: 460 (2002).
883:
884: \bibitem{gaslambda} H. J. de Vega, J. A. Siebert,
885: Nucl. Phys. {\bf B 707}, 529 (2005).
886:
887: \bibitem{Emdsg} R. Emden, Gaskugeln, Teubner, Leipzig und Berlin, 1907.
888:
889: \bibitem{Chandrsg} S. Chandrasekhar, `An introduction to the Study
890: of Stellar Structure', Chicago Univ. Press, 1939.
891:
892: \bibitem{Ebertsg} R. Ebert, Z. Astrophys. {\bf 37}, 217 (1955).
893:
894: \bibitem{Bonnorsg} W.B.Bonnor, Mon. Not. R. astr. Soc. {\bf 116}, 351 (1956).
895:
896: \bibitem{Antonovsg} V. A. Antonov, Vest. Leningrad Univ. 7, 135 (1962).
897:
898: \bibitem{Lyndsg} D. Lynden-Bell and R Wood,
899: Mon. Not. R. astr. Soc. {\bf 138}, 495 (1968).
900:
901: \bibitem{Katzsg} G. Horwitz and J. Katz, Ap. J. {\bf 211}, 226 (1977) and
902: {\bf 222}, 941 (1978).
903:
904: \bibitem{Padmanabhansg} T. Padmanabhan, Phys. Rep. 188, 285 (1990).
905:
906: \bibitem{Saslawsg} W. C. Saslaw,
907: `Gravitational Physics of stellar and galactic systems',
908: Cambridge Univ. Press, 1987.
909:
910: \bibitem{Jeans} J. H. Jeans, Astronomy and cosmogony,
911: Cambridge University Press (1919)
912:
913: \bibitem{BT} G. Bertin, M. Trenti, ApJ, {\bf 584}, 729, (2003).
914:
915: \bibitem{MC} See, for example, K. Binder, D. W. Heerman, Monte-Carlo Simulations
916: in Statisical Physics, Springer Series in Solid State, Vol.80, Berlin, 1988.
917:
918:
919: \bibitem{vortex} H. J. de Vega and F. A. Schaposnik, Phys. Rev. {\bf
920: D14}, 1100 (1976). A. Gonz\'alez Arroyo, A. Ramos, JHEP 0407, 008 (2004).
921:
922: \bibitem{lam} S. Perlmutter et al., ApJ, {\bf 517}, 565 (1999).
923:
924: A. G. Riess et al., Astron. J, {\bf 116 } 1009 (1998).
925:
926: See for a review, B. P. Schmidt in Phase Transitions in the Early
927: Universe: Theory and Observations, NATO ASI, Edited by H J de Vega, I
928: M Khalatnikov and N. S\'anchez, Series II, vol. 40, Kluwer, 2001.
929:
930: \end{thebibliography}
931: \end{document}
932:
933: