cond-mat0106272/a.tex
1: \documentclass[12pt]{article}
2: \usepackage{graphicx}
3: \catcode`\@=11
4: \topmargin 0pt
5: \oddsidemargin 0pt
6: \headheight 0pt
7: \headsep 0pt
8: \textheight 9in
9: \textwidth 6.25in
10: \marginparwidth.875in
11: \def\numberbysection{\@addtoreset{equation}{section}
12: \def\theequation{\thesection.\arabic{equation}}}
13: \def\baselinestretch{1.1}
14: \numberbysection
15: \newcommand{\abs}[1]{\left\vert#1\right\vert}
16: \newcommand{\be}{\[}
17: \newcommand{\beq}{\begin{equation}}
18: \newcommand{\bea}{\begin{eqnarray*}}
19: \newcommand{\beqa}{\begin{eqnarray}}
20: \renewcommand{\d}{{{\rm d}}}
21: \newcommand{\ds}{\displaystyle}
22: \newcommand{\de}{^{(2)}}
23: \newcommand{\dpar}{\partial}
24: \newcommand{\e}{{\rm e}}
25: \newcommand{\eq}{{\rm eq}}
26: \newcommand{\ee}{\]}
27: \newcommand{\eeq}{\end{equation}}
28: \newcommand{\eea}{\end{eqnarray*}}
29: \newcommand{\eeqa}{\end{eqnarray}}
30: \newcommand{\frad}[2]{\displaystyle{\displaystyle#1\over\displaystyle#2}}
31: \newcommand{\fd}{fluc\-tu\-a\-tion-dis\-si\-pa\-tion }
32: \newcommand{\g}{\gamma}
33: \newcommand{\h}{h}
34: \newcommand{\mean}[1]{\left\langle#1\right\rangle}
35: \renewcommand{\i}{{\rm i}}
36: \renewcommand{\l}{\ell}
37: \newcommand{\p}{\varphi}
38: \newcommand{\pp}{\pi}
39: \newcommand{\re}{\mathop{\rm Re}}
40: \newcommand{\Var}{\mathop{\rm Var}}
41: \newcommand{\un}{^{(1)}}
42: \newcommand{\unde}{^{(1,2)}}
43: \newcommand{\vb}{{\vphantom{M}}}
44: \newcommand{\C}{{\cal C}}
45: \newcommand{\D}{{\cal D}}
46: \newcommand{\Fd}{Fluc\-tu\-a\-tion-dis\-si\-pa\-tion }
47: \renewcommand{\H}{{\cal H}}
48: \newcommand{\J}{J_{(\beta+1)/2}}
49: \renewcommand{\L}{L^{(\beta+1)/2}}
50: \renewcommand{\P}{{\cal P}}
51: \newcommand{\X}{{\cal X}}
52: \newcommand{\text}[1]{\textrm{#1}}
53: \def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}
54: \def\binom#1#2{{#1\choose #2}}
55: 
56: \begin{document}
57: \centerline{\Large\bf Nonequilibrium dynamics of the zeta urn model}
58: \vspace{1cm}
59: \centerline{\large
60: by C.~Godr\`eche$^{a,}$\footnote{godreche@spec.saclay.cea.fr}
61: and J.M.~Luck$^{b,}$\footnote{luck@spht.saclay.cea.fr}}
62: \vspace{1cm}
63: \centerline{$^a$Service de Physique de l'\'Etat Condens\'e,
64: CEA Saclay, 91191 Gif-sur-Yvette cedex, France}
65: \vspace{.1cm}
66: \centerline{$^b$Service de Physique Th\'eorique\footnote{URA 2306 of CNRS},
67: CEA Saclay, 91191 Gif-sur-Yvette cedex, France}
68: \vspace{1cm}
69: \begin{abstract}
70: We consider a mean-field dynamical urn model, defined by rules
71: which give the rate at which
72: a ball is drawn from an urn and put in another one,
73: chosen amongst an assembly.
74: At equilibrium, this model possesses
75: a fluid and a condensed phase, separated by a critical line.
76: We present an analytical study of the nonequilibrium properties of
77: the fluctuating number of balls in a given urn,
78: considering successively the temporal evolution of its distribution,
79: of its two-time correlation and response functions,
80: and of the associated \fd ratio,
81: both along the critical line and in the condensed phase.
82: For well separated times the \fd ratio admits non-trivial limit values,
83: both at criticality and in the condensed phase,
84: which are universal quantities depending continuously on temperature.
85: \end{abstract}
86: \vfill
87: %\noindent To be submitted for publication to the European Physical Journal B.
88: %\hfill S/01/000
89: \vskip -3pt
90: \noindent P.A.C.S.: 02.50.Ey, 05.40.+j, 61.43.Fs.
91: %\hfill T/01/065
92: \newpage
93: \section{Background: on dynamical urn models}
94: 
95: Dynamical urn models are simplified models of physical reality, which have
96: always played an important role in the elucidation of conceptual problems of
97: statistical mechanics and probability theory.
98: 
99: The ancestor and prototype of this class of models is the
100: Ehrenfest urn model~\cite{ehr}.
101: It was devised by P.~and T.~Ehrenfest, in their attempt to critically review
102: Boltzmann's $H-$theorem.
103: Consider $N$ balls, labeled from 1 to $N$, which
104: are distributed in two urns (or boxes).
105: At each time step a ball is chosen
106: at random (i.e., an integer between 1 and~$N$ is chosen at random), and
107: moved from the box in which it is, to the other box.
108: Let $N_{1}$ (respectively, $N_{2}$) be the numbers of balls in box number 1
109: (respectively, number~2) ($N_{1}+N_{2}=N$).
110: If the process is repeated
111: indefinitely, for any initial condition the system will relax to
112: equilibrium, characterized by a binomial distribution of balls in, say, box
113: number 1:
114: \begin{equation}
115: f_{k,\eq}=\P(N_{1}=k)=\binom{N}{k}\frac{1}{2^{N}}.
116: \label{fk2}
117: \end{equation}
118: This result is both intuitively translucent, and easy to derive (see below).
119: The partition function of the system is equal to (for $2$ boxes and $N$ balls)
120: \beq
121: Z(2,N)=\frac{2^{N}}{N!}.
122: \label{z2n}
123: \eeq
124: There are $2^N$ possibilities of distributing the balls amongst the 2 boxes,
125: however all $N!$ labelings of the balls are equivalent.
126: 
127: Finding the distribution of balls after $n$ steps,
128: $f_{k}(n)=\P(N_{1}(n)=k)$, requires more effort.
129: Kohlrausch and Schr\"{o}dinger~\cite{ks} found
130: the master equation for $f_{k}(n)$, which they interpreted as the
131: probability distribution of the position at time $n$ of a random walker
132: (played here by $N_{1}(n)$).
133: The full
134: solution of this master equation was given later on by Kac, Siegert and Hess
135: \cite{kac,kac2,sieg,hess}.
136: 
137: The Ehrenfest model is at the origin of a whole class of dynamical urn models,
138: which we name the \emph{Ehrenfest class}.
139: They generalize the original
140: Ehrenfest model in two ways, which we detail successively.
141: 
142: The first generalization consists in considering $M$ boxes instead of two.
143: Then, at equilibrium, the joint distribution of the occupation numbers
144: $N_{1}$, $N_{2}$, $\dots$, $N_{M}$, with
145: \[
146: \sum_{i=1}^{M}N_{i}=N,
147: \]
148: is the multinomial distribution
149: \[
150: \P(N_{1}=k_{1},\dots,N_{M}=k_{M})=\binom{N}{k_{1}\dots k_{M}}
151: \frac{1}{M^{N}},
152: \]
153: as a simple reasoning shows.
154: The marginal distribution of $N_{1}$ is
155: obtained by summation upon the other variables, and reads
156: \[
157: f_{k,\eq}=\P(N_{1}=k)=\binom{N}{k}\frac{1}{M^{N}},
158: \]
159: which is a simple generalization of~(\ref{fk2}).
160: In the thermodynamic
161: limit $N\rightarrow\infty $, $M\rightarrow\infty $, with fixed density
162: $\rho=N/M$, this yields a Poisson law:
163: \[
164: f_{k,\eq}=\e^{-\rho }\frac{\rho ^{k}}{k!}.
165: \]
166: The partition function of the system is equal to (for $M$ boxes and $N$
167: balls)
168: \begin{equation}
169: Z(M,N)=\frac{M^{N}}{N!},
170: \label{zinfty}
171: \end{equation}
172: as a simple generalization of the reasoning leading to~(\ref{z2n}) shows.
173: 
174: The dynamics of the Ehrenfest model and of its generalization to $M$ boxes
175: takes
176: place at infinite temperature because there is no constraint on the move of
177: the drawn ball.
178: The second direction of generalization consists in defining
179: these models at finite temperature, by introducing energy.
180: We assume that the energy is a sum of contributions of independent boxes:
181: \[
182: E(N_{1},\dots,N_{M})=\sum_{i=1}^{M}E(N_{i}),
183: \]
184: and choose a rule obeying detailed balance for the move of the drawn ball.
185: For instance, for the Metropolis rule, the move is
186: allowed with probability $\min(1,\exp(-\beta\Delta E))$.
187: Heat-bath dynamics is another possible choice (see below).
188: 
189: The fundamental ingredients for the definition of the models belonging to
190: the Ehrenfest class are therefore
191: 
192: \begin{description}
193: \item [(i)] the statistics: a ball is chosen at random, and put in a box
194: chosen at random,
195: 
196: \item [(ii)] the choice of
197: the energy of a box $E(N_{i})$, and of a dynamical rule (Metropolis
198: or heat-bath),
199: 
200: \item [(iii)] the geometry: for instance, boxes may be ordered on a line, or
201: on the contrary be all connected.
202: For short, we designate the latter geometry as the mean-field case.
203: \end{description}
204: 
205: 
206: The backgammon model~\cite{ritort} is a representative of the Ehrenfest class,
207: corresponding to the choice (where $\delta$ is Kronecker symbol)
208: \begin{equation}
209: E(N_{i})=-\delta(N_{i},0).\label{back}
210: \end{equation}
211: This model has been extensively studied, mainly in its mean-field
212: formulation~\cite{fr,gbm,gl}.
213: 
214: The dynamical (and equilibrium) properties of the Ehrenfest class depend
215: crucially on the choice of statistics described in (i), which we will
216: briefly refer to as the \emph{ball-box statistics}.
217: However, other choices
218: are possible, which define new classes of dynamical urn models.
219: The class which we will refer to for short as the \emph{Monkey class},
220: because it corresponds to the image of a monkey playing at exchanging
221: balls between boxes, is defined by:
222: 
223: \begin{description}
224: \item [(i)] the statistics: a box is chosen at random, from which any ball is
225: drawn, and put in another box, chosen at random (\emph{box-box statistics}),
226: 
227: \item [(ii)] the choice of energy and dynamical rule (as above),
228: 
229: \item [(iii)] the geometry (as above).
230: \end{description}
231: 
232: A first example of a model belonging to this class corresponds to taking
233: definition~(\ref{back}) for the energy~\cite{gbm}.
234: This model, referred to as model B
235: in~\cite{gbm}, possesses non-trivial dynamical properties~\cite{gbm,gl}.
236: 
237: A second example, inspired from quantum gravity, corresponds to
238: taking~\cite{bia}
239: \begin{equation}
240: E(N_{i})=\ln(N_{i}+1).
241: \label{defE}
242: \end{equation}
243: It presents interesting properties both at equilibrium~\cite{bia} and in
244: nonequilibrium situations~\cite{dgc}.
245: In contrast with the backgammon model, or with
246: model B, it possesses a phase transition between a fluid phase and a
247: condensed phase at finite temperature~\cite{bia}.
248: We will refer to this model as the \emph{zeta urn model}, for reasons which
249: will appear clear in the sequel.
250: The present work is entirely devoted
251: to the study of the nonequilibrium behavior of the zeta urn model
252: in the mean-field geometry.
253: 
254: Before specializing to this model, let us first present, in parallel,
255: some formalism which applies to the two classes of models defined above,
256: in order to underline the fundamental role played by the choice of
257: statistics for both the equilibrium and nonequilibrium properties of the
258: models.
259: Note that the equilibrium properties of the dynamical urn models
260: defined above are independent of the geometry, because boxes are independent.
261: 
262: For the Ehrenfest class, the partition function reads
263: \begin{equation}
264: Z(M,N)=\sum_{N_{1}}\cdots\sum_{N_{M}}\frac{\,p_{N_{1}}}{N_{1}!}\cdots
265: \frac{p_{N_{M}}}{N_{M}!}\;\delta\left(\sum_{i}N_{i},N\right),
266: \label{jstar}
267: \end{equation}
268: where
269: $$
270: p_{N_{i}}=\e^{-\beta E(N_i)}
271: $$
272: is the unnormalized Boltzmann weight attached to box number $i$.
273: For the Monkey class we have
274: \beq
275: Z(M,N)=\sum_{N_{1}}\cdots\sum_{N_{M}}\,p_{N_{1}}\cdots p_{N_{M}}\;\delta
276: \left(\sum_{i}N_{i},N\right).
277: \label{Zbb}
278: \eeq
279: 
280: Using the integral representation $2\i\pi\delta(m,n)=\oint\d z\,z^{m-n-1}$,
281: we obtain
282: \begin{equation}
283: Z(M,N)=\oint\frac{\d z}{2\i\pi z^{N+1}}\left[P(z)\right]^{M},
284: \label{contour}
285: \end{equation}
286: where
287: \begin{eqnarray}
288: P(z)=\sum_{k=0}^\infty\frac{p_{k}}{k!}\,z^k{\hskip.75cm}
289: &&(\mathrm{Ehrenfest}),
290: \label{PzE}\\
291: =\sum_{k=0}^\infty p_{k}\,z^{k}\qquad&&(\mathrm{Monkey}).
292: \label{Pzbb}
293: \end{eqnarray}
294: The equilibrium properties of the models are therefore entirely encoded in
295: the tem\-pe\-ra\-ture-de\-pen\-dent generating series $P(z)$.
296: The presence or absence of the factorial term $k!$
297: has direct implication on the analytic structure of the series, and by
298: consequence on the possible existence of a phase transition at finite
299: temperature.
300: 
301: The equilibrium probability distribution of the occupation number
302: $N_{1}$ reads, for the Ehrenfest class,
303: \begin{eqnarray}
304: f_{k,\eq} &=&\P(N_{1}=k)=\left\langle\delta\left( N_{1},k\right)
305: \right\rangle
306: \nonumber\\
307: &=&\frac{1}{Z(M,N)}\sum_{N_{1}}\dots\sum_{N_{M}}\delta\left(
308: N_{1},k\right)\frac{\,p_{N_{1}}}{N_{1}!}\dots\frac{p_{N_{M}}}{N_{M}!}
309: \;\delta\left(\sum_{i}N_{i},N\right)
310: \nonumber\\
311: &=&\frac{p_{k}}{k!}\frac{Z(M-1,N-k)}{Z(M,N)}.
312: \label{fkE}
313: \end{eqnarray}
314: For the Monkey class one obtains
315: \begin{equation}
316: f_{k,\eq}=p_{k}\frac{Z(M-1,N-k)}{Z(M,N)}.
317: \label{fkbb}
318: \end{equation}
319: At infinite temperature, equation~(\ref{contour}),
320: together with~(\ref{PzE}) and~(\ref{Pzbb}),
321: respectively lead to~(\ref{zinfty}) and to
322: \begin{equation}
323: Z(M,N)=\frac{(M+N-1)!}{(M-1)!\,N!}.
324: \label{jmonkey}
325: \end{equation}
326: 
327: In the thermodynamic limit at fixed density $\rho$,
328: the free energy per box is defined as
329: \[
330: \beta F=-\lim_{M\to\infty}\frac{1}{M}\ln Z(M,N),\qquad N\approx M\rho.
331: \]
332: At infinite temperature, equations~(\ref{zinfty}) and~(\ref{jmonkey}) yield
333: \bea
334: \lim_{\beta\to0}\,\beta F
335: =\rho\ln\rho-\rho{\hskip 3.525cm}&&(\mathrm{Ehrenfest}),\\
336: =\rho\ln\rho-(\rho+1)\ln(\rho+1)\qquad&&(\mathrm{Monkey}).
337: \eea
338: At finite temperature, the free energy can be obtained by evaluating the
339: contour integral in~(\ref{contour}) by the saddle-point method.
340: The saddle-point value $z_{s}$ of $z$
341: is a function of temperature and density through the equation
342: \beq
343: \frac{z_{s}P^{\prime }(z_{s})}{P(z_{s})}=\rho,
344: \label{col}
345: \eeq
346: and the free energy per box reads
347: \beq
348: \beta F=\rho\ln z_{s}-\ln P(z_{s}).
349: \label{free}
350: \eeq
351: Similarly, we obtain the following expressions
352: for the equilibrium occupation probabilities in the thermodynamic limit
353: \begin{eqnarray}
354: f_{k,\eq}=\frac{p_{k}}{k!}\frac{z_{s}^{k}}{P(z_{s})}\qquad
355: &&(\mathrm{Ehrenfest}),\label{fkeq1}\\
356: =p_{k}\frac{z_{s}^{k}}{P(z_{s})}{\hskip.91cm}&&(\mathrm{Monkey}).
357: \label{fkeq2}
358: \end{eqnarray}
359: This formalism will be illustrated below on the zeta urn model,
360: studied in this work.
361: 
362: For both classes of models, the temporal evolution of the
363: occupation probability
364: \[
365: f_{k}(t)=\P(N_{1}(t)=k)
366: \]
367: is given by the master equation
368: 
369: \begin{equation}
370: \frac{\d f_{k}(t)}{\d t}=\sum_{\l=0}^\infty
371: \left(\pp_{k+1,\l}+\pp_{\l,k-1}-\pp_{k,\l}-\pp_{\l,k}\right),
372: \label{master1}
373: \end{equation}
374: where $\pp_{k,\l}$ denotes the contribution of a move from the departure box,
375: containing $k$ balls, to the arrival box, containing $\l$ balls.
376: Restricting our study to the mean-field case,
377: we have (for $k,\l\geq 0$)
378: \begin{eqnarray}
379: \pp_{k,\l}=kf_{k}f_{\l}W_{k,\l}(1-\delta_{k,0})\qquad&&(\mathrm{Ehrenfest}),
380: \nonumber\\
381: =f_{k}f_{\l}W_{k,\l}(1-\delta_{k,0}){\hskip 1.05cm}&&(\mathrm{Monkey}),
382: \label{pkl}
383: \end{eqnarray}
384: where the term $1-\delta_{k,0}$ accounts for the fact that
385: the departure box cannot be empty,
386: and where the acceptance rate $W_{k,\l}$ depends on the dynamics chosen.
387: With the Metropolis rule we have
388: \[
389: W_{k,\l}=\min\left(1,\frac{p_{k-1}p_\l}{p_kp_{\l+1}}\right),
390: \]
391: while the heat-bath rule leads to~\cite{dgc}
392: \beq
393: W_{k,\l}=\frac{p_{\l+1}}{p_{\l}}
394: \left(\sum_{\l=0}^{\infty}f_{\l}\frac{p_{\l+1}}{p_{\l}}\right)^{-1},
395: \label{heat}
396: \eeq
397: which only depends on the label $\l$ of the arrival box.
398: 
399: In all cases, equation~(\ref{master1}) can be seen as the master equation of
400: a random walk for $N_{1}$, i.e., over the positive integers $k=0,1,\dots$,
401: \begin{eqnarray}
402: \frac{\d f_{k}(t)}{\d t} &=&\mu_{k+1}\,f_{k+1}+\lambda_{k-1}\,f_{k-1}-\left(
403: \mu_{k}+\lambda_{k}\right) f_{k}\qquad(k\ge1),
404: \label{master2a}\\
405: \frac{\d f_{0}(t)}{\d t} &=&\mu_{1}\,f_{1}-\lambda_{0}f_{0},
406: \label{master2b}
407: \end{eqnarray}
408: generalizing the result of Kohlrausch and
409: Schr\"{o}dinger for the Ehrenfest model.
410: In these equations,
411: \[
412: \lambda_k=\sum_{\l=0}^\infty\frac{\pp_{\l,k}}{f_k},\qquad
413: \mu_k=\sum_{\l=0}^\infty\frac{\pp_{k,\l}}{f_k}
414: \]
415: are, respectively, the hopping rate to the right,
416: corresponding to $N_{1}=k\rightarrow N_{1}=k+1$,
417: and to the left, corresponding to $N_{1}=k\rightarrow N_{1}=k-1$.
418: The equation for $f_{0}$ is special because one cannot select an empty box
419: as a departure box, nor can $N_1$ be negative, hence $\lambda_{-1}=\mu_0=0$.
420: In other words a partially absorbing barrier is present at site $k=0$.
421: The random walk is locally biased, to the right or to the left, according to
422: whether its velocity $\lambda_k-\mu_k$ is positive or negative, respectively.
423: It is easy to verify that~(\ref{master2a}) and~(\ref{master2b})
424: preserve the sum rules
425: \begin{eqnarray}
426: \sum_{k}f_{k}(t) &=&1,\label{sum1}\\
427: \sum_{k}k\,f_{k}(t) &=&\left\langle N_{1}(t)\right\rangle =\rho
428: \label{sum2},
429: \end{eqnarray}
430: expressing respectively the conservation of probability
431: and of the number of particles.
432: 
433: The equilibrium occupation probabilities~(\ref{fkeq1}),~(\ref{fkeq2})
434: are recovered as the unique stationary state
435: ($\d f_k/\d t=0$) of the master equations~(\ref{master2a}),~(\ref{master2b}).
436: In contrast, except at infinite temperature, where $W_{k,\l}=1$,
437: so that the rates $\lambda_{k}$ and $\mu_{k}$ simplify,
438: the master equations cannot be solved explicitly.
439: The difficulty comes from the fact that
440: the rates $\lambda_{k}$ and $\mu_{k}$ are functions of the $f_{k}$,
441: hence the master equations are non-linear.
442: However, as will be illustrated by the present work,
443: the long-time behavior of these equations is amenable to analytic computations.
444: 
445: To summarize, for the two classes of dynamical urn models described above,
446: finding the properties of their equilibrium states is in general easy, but
447: attaining dynamical properties is much more difficult to achieve.
448: However if
449: one restricts the study to the fluctuating number of balls in a given box,
450: denoted by $N_{1}(t)$ all throughout this paper, it is,
451: in some cases, possible to predict
452: the long-time behavior of its probability distribution, $f_{k}(t)$, and also
453: of its two-time correlation and response functions.
454: This has been done in
455: previous studies for the backgammon model~\cite{fr,gbm,gl}.
456: 
457: In the present work we focus our interest on the mean-field
458: dynamical urn model
459: defined by the choice of energy~(\ref{defE}), and box-box statistics
460: --the zeta urn model.
461: We pursue the investigation of its
462: nonequilibrium properties, initiated in~\cite{dgc}.
463: The next section is devoted to a
464: more complete presentation of the model and to an outline of this paper.
465: 
466: \section{The zeta urn model}
467: 
468: At equilibrium, the zeta urn model is defined
469: by its partition function~(\ref{Zbb}), where the Boltzmann weight
470: \beq
471: p_{N_i}=\e^{-\beta E(N_i)}=(N_i+1)^{-\beta}
472: \label{boltz}
473: \eeq
474: corresponds to the choice of energy~(\ref{defE})~\cite{bia}.
475: The equilibrium phase diagram of the model~\cite{bia}
476: easily follows from the analysis of the previous section
477: (see~(\ref{contour})--(\ref{fkeq2})).
478: 
479: At low enough temperature $(\beta>2)$, there is a finite critical density:
480: \be
481: \rho_c=\frac{P'(1)}{P(1)}
482: =\frac{\zeta(\beta-1)-\zeta(\beta)}{\zeta(\beta)},
483: \ee
484: where $\zeta$ denotes Riemann's zeta function.
485: 
486: In the fluid phase $(\rho<\rho_c)$,
487: the equilibrium distribution~(\ref{fkeq2}) decays exponentially, since $z_s<1$.
488: 
489: At the critical density $(\rho=\rho_c)$, we have $z_s=1$, hence
490: \beq
491: f_{k,\eq}=\frac{p_k}{P(1)}=\frac{(k+1)^{-\beta}}{\zeta(\beta)},
492: \label{fkc}
493: \eeq
494: which is known as the zeta distribution.
495: In the regular part of the critical line~($\beta>3$),
496: the mean squared population is finite, and equal to
497: \be
498: \mean{N_i^2}=\sum_{k=0}^\infty k^2\,f_{k,\eq}
499: =\mu_c=\frac{\zeta(\beta-2)-2\zeta(\beta-1)+\zeta(\beta)}{\zeta(\beta)},
500: \ee
501: while it is infinite in the strong-fluctuation case~($2<\beta<3$).
502: Throughout the following, we shall restrict the study
503: to the regular part of the critical line.
504: 
505: In the condensed phase $(\rho>\rho_c)$,
506: a macroscopic condensate of particles appears.
507: Indeed, equation~(\ref{fkc}) still applies to all the boxes but one,
508: in which an extensive number of particles,
509: of order $N-M\rho_c=M(\rho-\rho_c)$, is condensed.
510: 
511: The dynamical definition of the model was given in the previous section.
512: For heat-bath dynamics, equation~(\ref{master1}),
513: together with~(\ref{pkl}) and~(\ref{heat}), leads to~\cite{dgc}
514: \beqa
515: &&\frad{\d f_k(t)}{\d t}=f_{k+1}(t)+\sigma(t)r_{k-1}f_{k-1}(t)
516: -\left(1+\sigma(t)r_k\right)f_k(t)\qquad(k\ge1),
517: \nonumber\\
518: &&\frad{\d f_0(t)}{\d t}=f_1(t)-\sigma(t)r_0f_0(t),
519: \label{df}
520: \eeqa
521: with
522: \be
523: r_k=\frac{p_{k+1}}{p_k}=\frac{f_{k+1,\eq}}{f_{k,\eq}}
524: =\left(\frac{k+1}{k+2}\right)^\beta
525: \ee
526: and
527: \be
528: \sigma(t)=\frad{1-f_0(t)}{\sum_{k=0}^\infty r_kf_k(t)}.
529: \ee
530: We assume that at time $t=0$ the system is quenched
531: from its infinite-temperature equilibrium state
532: to a finite temperature $T=1/\beta$.
533: Hence, by~(\ref{col}) and~(\ref{fkeq2}),
534: the initial occupation probabilities read
535: \beq
536: f_k(0)=\frac{\rho^k}{(1+\rho)^{k+1}}.
537: \label{fk0}
538: \eeq
539: 
540: In the fluid phase~($\rho<\rho_c$),
541: the equilibrium distribution $f_{k,\eq}$~(\ref{fkeq2})
542: is a stationary solution of~(\ref{df}), corresponding to $\sigma_\eq=z_s$.
543: The convergence of $f_k(t)$
544: towards $f_{k,\eq}$ is characterized by a finite relaxation time,
545: depending on $\beta$ and $\rho$.
546: 
547: The long-time behavior of the distribution $f_k(t)$,
548: both at criticality~($\rho=\rho_c$) and in the condensed phase~($\rho>\rho_c$),
549: is the subject of the next section.
550: In section 4 we establish the dynamical equations obeyed by the
551: two-time correlation and response functions of the population of a given box.
552: It is indeed well-known that
553: nonequilibrium properties are more fully revealed by
554: two-time observables~\cite{ck,aging}.
555: Section 5 is devoted to the analysis of the equilibrium properties of
556: these functions at criticality.
557: Section 6 is devoted to the analysis of their nonequilibrium properties at
558: criticality, including the violation of the fluctuation-dissipation theorem.
559: In section 7 we briefly investigate the nonequilibrium properties of the model
560: in the condensed phase.
561: 
562: \section{Long-time behavior of occupation probabilities}
563: 
564: \subsection{At criticality~($\rho=\rho_c$)}
565: 
566: We investigate how, starting from the disordered initial condition~(\ref{fk0}),
567: the occupation probabilities $f_k(t)$ converge toward
568: their critical equilibrium values $f_{k,\eq}$, given by~(\ref{fkc}),
569: which are the stationary solutions of equations~(\ref{df})
570: corresponding to $\sigma_\eq=1$.
571: In analogy with the analysis done in~\cite{dgc},
572: we anticipate that $\sigma(t)$ converges to this value as a power law:
573: \be
574: \sigma(t)\approx 1+A\,t^{-\omega},
575: \ee
576: and consider two regimes:
577: 
578: \medskip
579: \noindent{\bf Regime I}: $k$ fixed and $t\gg1$
580: 
581: This is the ``short-distance'' regime,
582: considering $k$ as the position of a fictitious random walker, as in section 1.
583: It is therefore analogous to the Porod regime
584: for phase-ordering systems~\cite{bray,glglau}.
585: 
586: Setting
587: \beq
588: f_k(t)\approx f_{k,\eq}\left(1+v_k\,t^{-\omega}\right),
589: \label{f1}
590: \eeq
591: equation~(\ref{df}) yields
592: \beq
593: v_k=v_0+A k,
594: \label{q1}
595: \eeq
596: where $v_0$ and $A$ are determined below.
597: 
598: \medskip
599: \noindent{\bf Regime II}: $k$ and $t$ simultaneously large (scaling regime)
600: 
601: In this regime, we look for a similarity solution
602: to~(\ref{df}) of the form
603: \beq
604: f_k(t)\approx f_{k,\eq}\,F(u),\qquad u=k\,t^{-1/2}.
605: \label{fsca}
606: \eeq
607: The structure of the master equations~(\ref{df})
608: indeed dictates that the scaling variable is the combination $kt^{-1/2}$.
609: Starting from a random initial condition, for a large but finite time~$t$,
610: and for $k$ much smaller than an ordering size of order $t^{1/2}$,
611: the system looks critical, i.e.,
612: the distribution $f_k(t)$ has essentially converged toward
613: the equilibrium distribution $f_{k,\eq}$.
614: This implies $F(0)=1$.
615: To the contrary, for $k\gg t^{1/2}$, the system still looks disordered,
616: i.e., the $f_k(t)$ fall off very fast.
617: Hence $F(\infty)=0$.
618: It will indeed be shown below that $f_k(t)\sim\exp(-k^2/(4t))$,
619: as the scaling function falls of very fast for $u\gg1$: $F(u)\sim\exp(-u^2/4)$.
620: Note the close analogy between the present situation and critical coarsening
621: for ferromagnetic spin systems~\cite{bray,glcrit},
622: where the scaling variable is $\abs{{\bf r}}t^{-1/z}$,
623: and where the role of $f_k(t)$ is played by the equal-time correlation
624: function $C(\abs{{\bf r}},t)$.
625: 
626: In order to determine the exponent $\omega$, we use
627: the sum rules~(\ref{sum1}) and~(\ref{sum2}), which yield respectively
628: \beqa
629: &&t^{-\omega}(v_0+A\rho_c)
630: =t^{-(\beta-1)/2}I_1,\qquad
631: I_1=\frac{1}{\zeta(\beta)}
632: \int_0^\infty u^{-\beta}\left(1-F(u)\right)\d u,\label{sumf1}\\
633: &&t^{-\omega}(v_0\rho_c+A\mu_c)
634: =t^{-(\beta-2)/2}I_2,\qquad
635: I_2=\frac{1}{\zeta(\beta)}
636: \int_0^\infty u^{1-\beta}\left(1-F(u)\right)\d u.\label{sumf2}
637: \eeqa
638: These equations are compatible only if the right-hand side
639: of equation~(\ref{sumf1}) is subleading.
640: In the case of a regular critical point ($\beta>3$), we thus obtain
641: \be
642: \omega=\frac{\beta-2}{2}>\frac{1}{2}
643: \ee
644: and
645: \be
646: A=\frac{I_2}{\mu_c-\rho_c^2},\qquad v_0=-\frac{\rho_c\,I_2}{\mu_c-\rho_c^2}.
647: \ee
648: 
649: Inserting the form~(\ref{fsca}) into~(\ref{df}),
650: and using the fact that $\omega>1/2$, leads to the differential equation
651: \beq
652: \D F(u)=0,
653: \label{fdif}
654: \eeq
655: where $\D$ is the linear differential operator
656: \beq
657: \D=-\frac{\d^2}{\d u^2}
658: +\left(-\frac{u}{2}+\frac{\beta}{u}\right)\frac{\d}{\d u}.
659: \label{ddef}
660: \eeq
661: The solution of~(\ref{fdif}) is
662: \beq
663: F(u)=\frac{2^{-\beta}}{\Gamma\!\left(\frac{\beta+1}{2}\right)}
664: \int_u^\infty y^\beta\,\e^{-y^2/4}\,\d y.
665: \label{fex}
666: \eeq
667: 
668: We present in Appendix A an alternative way of solving equation~(\ref{fdif}),
669: using the Mellin transformation.
670: In particular~(\ref{melumf}) yields the explicit expression
671: \be
672: I_2=\frac{M_{1-F}(\beta-2)}{\zeta(\beta)}=\frac{\pi^{1/2}\,2^{1-\beta}}
673: {(\beta-2)\,\Gamma\!\left(\frac{\beta+1}{2}\right)
674: \zeta(\beta)}.
675: \ee
676: 
677: As an illustration of the above,
678: let us determine how the variance of the population of box number 1,
679: $\Var N_1(t)=\mean{N_1(t)^2}-\mean{N_1(t)}^2=\mean{N_1(t)^2}-\rho_c^2$,
680: converges at long times to its equilibrium value $\mu_c-\rho_c^2$.
681: Only the scaling regime matters for the long-time behavior of this quantity,
682: and of all the quantities at criticality to be considered hereafter.
683: Using~(\ref{fsca}), we obtain
684: \be
685: \Var N_1(t)-(\mu_c-\rho_c^2)\approx-\Delta\,t^{-(\beta-3)/2},
686: \ee
687: with, by~(\ref{melumf}),
688: \be
689: \Delta=\frac{M_{1-F}(\beta-3)}{\zeta(\beta)}=\frac{2^{3-\beta}}
690: {(\beta-3)\,\Gamma\!\left(\frac{\beta+1}{2}\right)\zeta(\beta)}.
691: \ee
692: 
693: \subsection{In the condensed phase~($\rho>\rho_c$)}
694: 
695: We set
696: \be
697: \sigma(t)\approx 1+A\,t^{-1/2},
698: \ee
699: and consider the same two regimes as at criticality
700: (see~\cite{dgc} for more details).
701: 
702: \medskip
703: \noindent{\bf Regime~I}: $k$ fixed and $t\gg1$ (short-distance regime)
704: 
705: Equations~(\ref{f1}) and~(\ref{q1}) still hold,
706: but now with $\omega=1/2$, for any $\beta>2$.
707: 
708: \medskip
709: \noindent{\bf Regime~II}: $k$ and $t$ simultaneously large (scaling regime)
710: 
711: Looking for a similarity solution of equations~(\ref{df}), of the form
712: \beq
713: f_k(t)\approx\frac{F(u)}{t},\qquad u=k\,t^{-1/2},
714: \label{cofsca}
715: \eeq
716: we obtain for the scaling function $F(u)$ the linear differential
717: equation~\cite{dgc}
718: \beq
719: \frac{\d^2F}{\d u^2}
720: +\left(\frac{u}{2}-A+\frac{\beta}{u}\right)\frac{\d F}{\d u}
721: +\left(1-\frac{\beta}{u^2}\right)F=0.
722: \label{cofdif}
723: \eeq
724: 
725: Following~\cite{dgc}, we notice that the amplitude $A$
726: is determined by the fact that equation~(\ref{cofdif}) has an acceptable
727: solution $F(u)$, vanishing as $u\to0$ and $u\to\infty$.
728: The normalization of the solution $F(u)$ is determined by
729: the sum rule~(\ref{sum2}), which yields
730: \beq
731: \int_0^\infty u\,F(u)\,\d u=\rho-\rho_c.
732: \label{conorma}
733: \eeq
734: The parameter $v_0$ entering Regime~I is determined by the
735: sum rule~(\ref{sum1}), leading to
736: \be
737: v_0=-A\,\rho_c-\int_0^\infty F(u)\,\d u.
738: \ee
739: 
740: At variance with equation~(\ref{fdif}),
741: the differential equation~(\ref{cofdif}) cannot be solved in closed form.
742: It can be recast in Schwarzian form, without first derivative, by setting
743: \be
744: F(u)=u\,Y(u)\,H(u),
745: \ee
746: with
747: \beq
748: Y(u)=u^{-\beta/2-1}\,\e^{Au/2-u^2/8}.
749: \label{coeta}
750: \eeq
751: We thus obtain for $H(u)$ a differential equation of the form
752: \beq
753: \H\,H=0,
754: \label{cohdif}
755: \eeq
756: with
757: \beq
758: \H=-\frac{\d^2}{\d u^2}+W(u),
759: \label{cohdef}
760: \eeq
761: and where the potential $W(u)$ reads
762: \beq
763: W(u)=\frac{u^2}{16}-\frac{Au}{4}+\frac{\beta-3+A^2}{4}
764: -\frac{A\beta}{2u}+\frac{\beta(\beta+2)}{4u^2}.
765: \label{cow}
766: \eeq
767: 
768: Equation~(\ref{cohdif}) is a biconfluent Heun equation~\cite{heun}.
769: We will therefore refer to $\H$ as the Heun operator,
770: denoting its discrete eigenvalues by $E_n$
771: and the corresponding eigenfunctions bu $H_n(u)$.
772: Equation~(\ref{cohdif}) implies that the ground-state eigenvalue reads $E_0=0$,
773: while the associated eigenfunction $H_0(u)$
774: is simply related to the scaling function~$F(u)$:
775: \beq
776: F(u)=c\,u\,Y(u)\,H_0(u).
777: \label{coftoh}
778: \eeq
779: By~(\ref{conorma}) we have
780: \beq
781: c=\frad{\rho-\rho_c}{\int_0^{\infty^\vb}u^2\,Y(u)\,H_0(u)\,\d u}.
782: \label{coc}
783: \eeq
784: 
785: The spectrum of the operator $\H$,
786: and related quantities such as the scaling function $F(u)$,
787: can be further investigated in the limiting regimes of high and low
788: temperature.
789: 
790: \medskip
791: \noindent{\bf High temperature $(\beta\to0)$}
792: 
793: The analysis of this limiting situation will be helpful in section 7,
794: although it is of no direct physical relevance,
795: since the condensed phase only exists at low enough temperature $(\beta>2)$.
796: 
797: For $\beta=0$ and $A=0$, the potential~(\ref{cow}) becomes $W(u)=u^2/16-3/4$,
798: so that $\H$ is
799: the Hamiltonian of a harmonic oscillator on the half-line $u\ge0$,
800: up to a scale, with Dirichlet boundary condition at $u=0$.
801: Its spectrum is $E_n=n$ $(n=0,1,\dots)$,
802: and the (unnormalized) ground-state eigenfunction reads
803: \beq
804: H_0(u)=u\,\e^{-u^2/8},
805: \label{cohih}
806: \eeq
807: hence
808: \beq
809: F(u)=\frac{\rho-\rho_c}{2\,\pi^{1/2}}\,u\,\e^{-u^2/4}.
810: \eeq
811: 
812: Perturbation theory can then be used
813: to determine the small-$\beta$ behavior of the amplitude~$A$.
814: Expressing that the lowest eigenvalue reads $E_0=0$
815: with no correction, we obtain
816: \be
817: \int_0^\infty\left(-Au+\beta+\frac{2\beta}{u^2}\right)H_0^2(u)\,\d u+\cdots=0,
818: \ee
819: hence
820: \beq
821: A\approx\frac{\pi^{1/2}}{2}\beta.
822: \label{cohia}
823: \eeq
824: 
825: \medskip
826: \noindent{\bf Low temperature $(\beta\to\infty)$}
827: 
828: In this other limiting situation, the operator $\H$ simplifies as follows.
829: If we rescale $A$ and $u$ according to $A\sim u\sim\beta^{1/2}$,
830: all terms in the expression~(\ref{cow}) for the potential $W(u)$
831: scale as $\beta$.
832: In other words, $\beta$ plays the role of $1/\hbar$ in Quantum Mechanics,
833: and $\beta\to\infty$ is the semi-classical regime.
834: Expressing that the minimum of the potential is at zero yields
835: the first estimates $A\approx u\approx(2\beta)^{1/2}$.
836: 
837: A more refined analysis consists in expanding the potential $W(u)$
838: around its minimum.
839: Setting
840: \beq
841: A=(2\beta)^{1/2}-a,\qquad
842: u=(2\beta)^{1/2}+\beta^{1/8}v,
843: \label{coloset}
844: \eeq
845: the operator $\H$ becomes
846: \beq
847: \H\approx\frac{a^2-2}{4}
848: +\beta^{-1/4}\left(-\frac{\d^2}{\d v^2}+\frac{a}{2^{5/2}}v^2\right).
849: \label{cohred}
850: \eeq
851: The expression in the parentheses is again proportional
852: to the Hamiltonian of a harmonic oscillator.
853: Expressing that the lowest eigenvalue of the right-hand-side of~(\ref{cohred})
854: is $E_0=0$ determines $a=2^{1/2}-2^{-1/2}\beta^{-1/4}$, hence
855: \be
856: A\approx(2\beta)^{1/2}-2^{1/2}+2^{-1/2}\beta^{-1/4}.
857: \ee
858: 
859: The spectrum of $\H$ reads $E_n\approx\beta^{-1/4}\,n$ ($n=0,1,\dots$),
860: and the (unnormalized) ground-state eigenfunction reads
861: $H_0(u)\approx\e^{-v^2/4}$, i.e.,
862: \beq
863: H_0(u)\approx\exp\left(-\frac{(u-(2\beta)^{1/2})^2}{4\beta^{1/4}}\right).
864: \label{coloh}
865: \eeq
866: Both this expression and the function $Y(u)$ become singular
867: in the $\beta\to\infty$ limit,
868: so that a direct analysis of equation~(\ref{cofdif}) is needed
869: in order to derive
870: the behavior of the scaling function $F(u)$ at low temperature.
871: The above analysis suggests to look for a solution
872: depending on the scaling variable $y=(2\beta)^{-1/2}u$.
873: The term involving the second-order derivative $\d^2F/\d u^2$
874: is then negligible.
875: The simplified form of equation~(\ref{cofdif}), namely
876: \[
877: y(y-1)^2\,\frac{\d F}{\d y}+(2y^2-1)F=0,\qquad y=(2\beta)^{-1/2}u,
878: \]
879: admits the normalized solution
880: \[
881: F(u)=\frac{\rho-\rho_c}{2E_1(1)\,\beta}
882: \,\frac{y}{(1-y)^3}\,\exp\left(-\frac{1}{1-y}\right)\qquad(0<y<1),
883: \]
884: where $E_1$ is the first exponential integral $(E_1(1)=0.219383934)$.
885: The scaling function $F(u)$ is therefore nonzero in the $\beta\to\infty$ limit
886: only for $y<1$, i.e., $u<(2\beta)^{1/2}$.
887: The upper bound $y=1$, i.e., $u=(2\beta)^{1/2}$,
888: coincides with the point where the eigenfunction $H_0(u)$,
889: as given by equation~(\ref{coloh}), peaks at low temperature.
890: 
891: \section{Two-time observables: dynamical equations}
892: 
893: In this section we will successively establish dynamical equations for
894: the two-time correlation function $C(t,s)$,
895: for its derivative $\dpar C(t,s)/\dpar s$,
896: and for the response function $R(t,s)$.
897: (See~\cite{gl} for similar techniques.)
898: 
899: We consider the (connected) two-time correlation function $C(t,s)$
900: between the population of box number 1 at times
901: $s$~(waiting time) and $t$~(observation time), with $0\le s\le t$:
902: \be
903: C(t,s)=\mean{N_1(t)N_1(s)}-\mean{N_1(t)}\mean{N_1(s)}
904: =\mean{N_1(t)N_1(s)}-\rho^2.
905: \ee
906: This definition can be recast as~\cite{gl}
907: \be
908: C(t,s)=\sum_{k=1}^\infty k\,\g_k(t,s)-\rho^2,
909: \ee
910: with
911: \be
912: \g_k(t,s)=\sum_{j=1}^\infty j\,f_j(s)\,\P\{N_1(t)=k\mid N_1(s)=j\}.
913: \ee
914: The evolution of the $\g_k(t,s)$ with respect to $t$ is given, for $t\ge s$,
915: by a master equation similar to~(\ref{df}):
916: \beqa
917: &&\frad{\dpar\g_k(t,s)}{\dpar t}=\g_{k+1}(t,s)+\sigma(t)r_{k-1}\g_{k-1}(t,s)
918: -\left(1+\sigma(t)r_k\right)\g_k(t,s)\qquad(k\ge1),
919: \nonumber\\
920: &&\frad{\dpar\g_0(t,s)}{\dpar t}=\g_1(t,s)-\sigma(t)r_0\g_0(t,s).
921: \label{dg}
922: \eeqa
923: These equations preserve the sum rule $\sum_{k}\g_k(t,s)=\rho$.
924: At $t=s$, the initial conditions are $\g_k(s,s)=k\,f_k(s)$,
925: implying
926: $$
927: C(s,s)=\sum_{k=1}^\infty k^2\,f_k(s)-\rho^2,
928: $$
929: which is the variance of the population of box number 1
930: at time $s$, as it should.
931: 
932: In the discussion of the \fd theorem, we will
933: need expressions of the time derivative $\dpar C(t,s)/\dpar s$.
934: We have
935: \be
936: \frac{\dpar C(t,s)}{\dpar s}=\sum_{k=1}^\infty k\,\p_k(t,s),
937: \ee
938: with
939: \be
940: \p_k(t,s)=\frac{\dpar\g_k(t,s)}{\dpar s}.
941: \ee
942: 
943: The evolution of the $\p_k(t,s)$ with respect to $t$ is again given,
944: for $t\ge s$, by equations similar to~(\ref{df}):
945: \beqa
946: &&\frad{\dpar\p_k(t,s)}{\dpar t}=\p_{k+1}(t,s)+\sigma(t)r_{k-1}\p_{k-1}(t,s)
947: -\left(1+\sigma(t)r_k\right)\p_k(t,s)\qquad(k\ge1),
948: \nonumber\\
949: &&\frad{\dpar\p_0(t,s)}{\dpar t}=\p_1(t,s)-\sigma(t)r_0\p_0(t,s),
950: \label{dp}
951: \eeqa
952: with initial conditions at $t=s$
953: \bea
954: &&\p_k(s,s)=-f_{k+1}(s)+\sigma(s)r_{k-1}f_{k-1}(s)\qquad(k\ge1),
955: \nonumber\\
956: &&\p_0(s,s)=-f_1(s),
957: \eea
958: so that
959: \be
960: \left(\frac{\dpar C(t,s)}{\dpar s}\right)_{t=s}
961: =\sigma(s)\sum_{k=1}^\infty k\,r_k\,f_k(s)+2(1-f_0(s))-\rho.
962: \ee
963: 
964: The two-time response function $R(t,s)$
965: is a measure of the change in the mean population of box number 1 at time $t$,
966: induced by an infinitesimal modulation of the conjugate variable,
967: i.e., the local chemical potential acting on the same box,
968: at the earlier time $s$.
969: 
970: In the presence of an arbitrary local, time-dependent chemical potential
971: $\mu(t)$, the energy of box number 1 at time $t$ reads
972: \be
973: E\left(N_1(t)\right)=\ln(N_1(t)+1)-\mu(t)N_1(t).
974: \ee
975: The occupation probabilities of this box now depend on $\mu(t)$:
976: we denote them by $f_k^\mu(t)$.
977: In the thermodynamic limit, i.e., to leading order as $M\to\infty$,
978: the occupation probabilities of all
979: the other boxes $(i=2,\dots,M)$ are still given by the $f_k(t)$.
980: 
981: The response function reads
982: \be
983: R(t,s)=\left(\frac{\delta\mean{N_1(t)}}{\delta\mu(s)}\right)_{\mu=0}
984: =\sum_{k=1}^\infty k\,\h_k(t,s),
985: \ee
986: with
987: \be
988: \h_k(t,s)=\left(\frac{\delta f^\mu_k(t)}{\delta\mu(s)}\right)_{\mu=0}.
989: \ee
990: 
991: The modified occupation probabilities $f_k^\mu(t)$ obey the dynamical equations
992: \beqa
993: &&\frad{\d f^\mu_k(t)}{\d t}
994: =f^\mu_{k+1}(t)+\sigma(t)\,\e^{\beta\mu(t)}r_{k-1}f^\mu_{k-1}(t)
995: -\left(1+\sigma(t)\,\e^{\beta\mu(t)}r_k\right)f^\mu_k(t)\qquad(k\ge1),
996: \nonumber\\
997: &&\frad{\d f^\mu_0(t)}{\d t}
998: =f^\mu_1(t)-\sigma(t)\,\e^{\beta\mu(t)}r_0f^\mu_0(t).
999: \label{dfa}
1000: \eeqa
1001: The initial values $f^\mu_k(0)=f_k(0)$ and the parameter $\sigma(t)$
1002: are unchanged.
1003: The dynamical equations~(\ref{dfa}) preserve the sum rule
1004: $\sum_k f^\mu_k(t)=1$.
1005: 
1006: Equations~(\ref{dfa}) imply that the $\h_k(t,s)$ obey, for $t>s$,
1007: \beqa
1008: &&\frad{\dpar\h_k(t,s)}{\dpar t}=\h_{k+1}(t,s)+\sigma(t)r_{k-1}\h_{k-1}(t,s)
1009: -\left(1+\sigma(t)r_k\right)\h_k(t,s)\qquad(k\ge1),
1010: \nonumber\\
1011: &&\frad{\dpar\h_0(t,s)}{\dpar t}=\h_1(t,s)-\sigma(t)r_0\h_0(t,s),
1012: \label{dh}
1013: \eeqa
1014: with initial conditions at $t=s$
1015: \bea
1016: &&\h_k(s,s)
1017: =\beta\sigma(s)\left(r_{k-1}f_{k-1}(s)-r_kf_k(s)\right)\qquad(k\ge1),
1018: \nonumber\\
1019: &&\h_0(s,s)=-\beta\sigma(s)r_0f_0(s),
1020: \eea
1021: so that
1022: \be
1023: R(s,s)=\beta\left(1-f_0(s)\right).
1024: \ee
1025: 
1026: The behavior of the two-time observables will now be successively investigated
1027: in the next three sections, first at criticality~($\rho=\rho_c$),
1028: both at equilibrium and in the nonequilibrium regime,
1029: and then in the condensed phase~($\rho>\rho_c$).
1030: 
1031: \section{Equilibrium critical dynamics}
1032: 
1033: When the waiting time $s$ becomes very large,
1034: keeping the difference $\tau=t-s\ge0$ fixed,
1035: two-time quantities reach their equilibrium values,
1036: which only depend on $\tau$, both in the fluid phase ($\rho<\rho_c$)
1037: and along the critical line ($\rho=\rho_c$).
1038: This section is devoted to the latter case.
1039: 
1040: The equilibrium correlation function $C_\eq(\tau)$ reads
1041: \be
1042: C_\eq(\tau)=\sum_{k=1}^\infty k\,\g_{k,\eq}(\tau)-\rho_c^2,
1043: \ee
1044: where the equilibrium values $\g_{k,\eq}(\tau)$ of the $\g_k(t,s)$ obey
1045: \beqa
1046: &&\frad{\d\g_{k,\eq}(\tau)}{\d t}
1047: =\g_{k+1,\eq}(\tau)+r_{k-1}\g_{k-1,\eq}(\tau)
1048: -\left(1+r_k\right)\g_{k,\eq}(\tau)
1049: \qquad(k\ge1),
1050: \nonumber\\
1051: &&\frad{\d\g_{0,\eq}(\tau)}{\d t}
1052: =\g_{1,\eq}(\tau)-r_0\g_{0,\eq}(\tau),
1053: \label{dgeq}
1054: \eeqa
1055: with initial conditions $\g_{k,\eq}(0)=k\,f_{k,\eq}$.
1056: This implies consistently that $C_\eq(0)=\mu_c-\rho_c^2$
1057: is the equilibrium variance of the population of a generic box.
1058: 
1059: As the time difference $\tau$ gets large,
1060: we have $\g_{k,\eq}(\tau)\to\rho_c\,f_{k,\eq}$, so that $C_\eq(\tau)\to0$.
1061: The decay of $C_\eq(\tau)$ for large $\tau$ can be investigated along the
1062: lines of section~3.1.
1063: The two regimes I and II are again to be considered separately,
1064: although results concerning the former will not be needed explicitly.
1065: In the scaling regime, we look for a similarity solution
1066: to equations~(\ref{dgeq}) of the form
1067: \beq
1068: \g_{k,\eq}(\tau)\approx
1069: f_{k,\eq}\,\tau^{1/2}\,G_\eq(u),\qquad u=k\,\tau^{-1/2},
1070: \label{geqsca}
1071: \eeq
1072: for which equations~(\ref{dgeq}) yield the differential equation
1073: \beq
1074: \left(\D+\frac12\right)G_\eq(u)=0,
1075: \label{geqdif}
1076: \eeq
1077: with boundary conditions $G_\eq(0)=0$ and $G_\eq(u)\approx u$ as $u\to\infty$.
1078: The differential operator $\D$ was defined in equation~(\ref{ddef}).
1079: 
1080: Equation~(\ref{geqdif}) can be solved by the method of Appendix A.
1081: With the definition~(\ref{mdef}), we have the functional equation
1082: \be
1083: \frac{M_{G_\eq}(z+2)}{M_{G_\eq}(z)}=\frac{z-1}{2(z+2)(\beta-1-z)},
1084: \ee
1085: whose suitably normalized solution reads
1086: \be
1087: M_{G_\eq}(z)=\frac
1088: {\pi^{1/2}\,\Gamma\!\left(\frac{z-1}{2}\right)\,
1089: \Gamma\!\left(\frac{\beta+1-z}{2}\right)}
1090: {2^{z+1}\,\Gamma\!\left(\frac{z+2}{2}\right)\,
1091: \Gamma\!\left(\frac{\beta}{2}\right)}
1092: \qquad(1<\re z<\beta+1).
1093: \ee
1094: 
1095: The decay of the equilibrium correlation function at large $\tau$
1096: is again dominated by the scaling regime, and the scaling form~(\ref{geqsca})
1097: yields
1098: \beq
1099: C_\eq(\tau)\approx A_\eq\,\tau^{-(\beta-3)/2},
1100: \label{ceq}
1101: \eeq
1102: with
1103: \be
1104: A_\eq=\frac{M_{G_\eq}(\beta-2)}{\zeta(\beta)}
1105: =\frac{\pi\,\Gamma\!\left(\frac{\beta-3}{2}\right)}
1106: {2^\beta\,\Gamma\!\left(\frac{\beta}{2}\right)^2\zeta(\beta)}.
1107: \ee
1108: 
1109: As expected, the \fd theorem holds at equilibrium.
1110: Indeed the equilibrium values $\p_{k,\eq}(\tau)$ and $\h_{k,\eq}(\tau)$
1111: of $\p_k(t,s)$ and $\h_k(t,s)$
1112: obey the same equations, identical to equations~(\ref{dgeq}),
1113: with initial conditions:
1114: \bea
1115: &&\h_{k,\eq}(0)=\beta\p_{k,\eq}(0)=
1116: \beta\left(f_{k,\eq}-f_{k+1,\eq}\right)\qquad(k\ge1),
1117: \nonumber\\
1118: &&\h_{0,\eq}(0)=\beta\p_{0,\eq}(0)=-\beta f_{1,\eq},
1119: \eea
1120: hence the identities
1121: \beq
1122: \h_{k,\eq}(\tau)=\beta\p_{k,\eq}(\tau)
1123: \eeq
1124: and
1125: \be
1126: R_\eq(\tau)=-\beta\,\frac{\d C_\eq(\tau)}{\d\tau}.
1127: \ee
1128: The last formula is the \fd theorem in its usual differential form.
1129: 
1130: \section{Nonequilibrium critical dynamics}
1131: 
1132: We now turn to the nonequilibrium critical behavior
1133: of the two-time correlation function $C(t,s)$,
1134: response function $R(t,s)$, and \fd ratio~\cite{ck,aging}
1135: \beq
1136: X(t,s)=\frac{R(t,s)}{\beta\,\frad{\dpar C(t,s)}{\dpar s}},
1137: \label{xdef}
1138: \eeq
1139: in the scaling regime where both time variables $s$ and $t$
1140: are large and comparable.
1141: Hereafter $x$ will denote the dimensionless time ratio
1142: \be
1143: x=\frac{t}{s}\ge1.
1144: \ee
1145: 
1146: We first analyze the scaling behavior of the correlation function $C(t,s)$.
1147: Looking for a two-variable scaling solution to equations~(\ref{dg})
1148: of the form
1149: \beq
1150: \g_k(t,s)\approx f_{k,\eq}\,t^{1/2}\,G(u,x),\qquad u=k\,t^{-1/2},\qquad x=t/s,
1151: \label{gsca}
1152: \eeq
1153: we obtain the partial differential equation
1154: \beq
1155: \left(x\frac{\dpar}{\dpar x}+\D+\frac12\right)G(u,x)=0
1156: \label{gdif}
1157: \eeq
1158: for the scaling function $G(u,x)$, with initial condition
1159: \beq
1160: G(u,1)=u F(u),
1161: \label{initg}
1162: \eeq
1163: and boundary condition $G(\infty,x)=0$ for all $x\ge1$.
1164: The function $F(u)$ is known from~(\ref{fex}),
1165: and the operator $\D$ is given by~(\ref{ddef}).
1166: 
1167: As shown in Appendix B, equation~(\ref{gdif})
1168: can be solved explicitly by the method of separation of variables.
1169: Equation~(\ref{gsca}) implies the scaling law
1170: \beq
1171: C(t,s)\approx s^{-(\beta-3)/2}\,\Phi(x),
1172: \label{csca}
1173: \eeq
1174: where
1175: \be
1176: \Phi(x)=\frac{x^{-(\beta-3)/2}}{\zeta(\beta)}\int_0^\infty
1177: u^{1-\beta}\,G(u,x)\,\d u.
1178: \ee
1179: Using~(\ref{gexpand}), we obtain
1180: \beq
1181: \Phi(x)=x^{-\beta/2}\sum_{n=0}^\infty A_n\,x^{-n},
1182: \label{phiexp}
1183: \eeq
1184: with
1185: \be
1186: A_n=\frac{a_n\,M_{G_n}(\beta-2)}{\zeta(\beta)}.
1187: \ee
1188: More explicitly, equations~(\ref{melgn}) and~(\ref{an}) imply that
1189: the leading coefficient of the expansion~(\ref{phiexp}) reads
1190: \beq
1191: A_0=\frac{\pi^{1/2}\,2^{3-\beta}\Gamma\!\left(\frac{\beta+4}{2}\right)}
1192: {3(\beta+1)\,\Gamma\!\left(\frac{\beta+1}{2}\right)^2\zeta(\beta)},
1193: \eeq
1194: while the coefficient ratios are rational functions of $\beta$:
1195: \beq
1196: \frac{A_n}{A_0}
1197: =\frac{3}{2^n}\prod_{j=0}^{n-1}(2j+\beta)\,\cdot\,
1198: \sum_{k=0}^n\frac{(-)^k}{(2k+3)k!(n-k)!}\prod_{\ell=0}^{k-1}
1199: \frac{2\ell+\beta+4}{2\ell+\beta+3},
1200: \eeq
1201: i.e.,
1202: \be
1203: \frac{A_1}{A_0}=\frac{\beta(2\beta+3)}{10(\beta+3)},\qquad
1204: \frac{A_2}{A_0}=\frac{\beta(\beta+2)(8\beta^2+52\beta+45)}
1205: {280(\beta+3)(\beta+5)},\qquad\hbox{etc.}
1206: \ee
1207: 
1208: The expansion~(\ref{phiexp})
1209: is convergent over the whole physical domain $(x>1)$.
1210: For $x\to1$, i.e., $\tau=t-s\ll s$, the equilibrium result~(\ref{ceq})
1211: is recovered as
1212: \beq
1213: \Phi(x)\approx A_\eq(x-1)^{-(\beta-3)/2}.
1214: \label{phieq}
1215: \eeq
1216: These properties can be checked by noticing that the expression~(\ref{lag})
1217: for the Laguerre polynomials simplifies to
1218: \be
1219: \L_n(u^2/4)\approx\left(\frac{2\,n^{1/2}}{u}\right)^{(\beta+1)/2}
1220: \J\left(u\,n^{1/2}\right),
1221: \ee
1222: where $J$ is the Bessel function,
1223: in the scaling regime where the order $n$ is large and $u$ is small.
1224: The subsequent integrals can be estimated for $n$ large, yielding
1225: \be
1226: A_n\approx
1227: \frac{\pi}{2^\beta\,\Gamma\!\left(\frac{\beta}{2}\right)^2\zeta(\beta)}
1228: \,n^{(\beta-5)/2}.
1229: \ee
1230: This asymptotic expression ensures the convergence
1231: of the series~(\ref{phiexp}) for all $x>1$,
1232: and establishes~(\ref{phieq}), including the prefactor.
1233: 
1234: We now turn to the derivative $\dpar C(t,s)/\dpar s$
1235: of the correlation function.
1236: Looking for a two-variable scaling solution to equations~(\ref{dp})
1237: in the scaling regime, of the form
1238: \beq
1239: \p_k(t,s)\approx f_{k,\eq}\,t^{-1/2}\,G\un(u,x),
1240: \label{g1sca}
1241: \eeq
1242: we get the partial differential equation
1243: \beq
1244: \left(x\frac{\dpar}{\dpar x}+\D-\frac12\right)G\un(u,x)=0,
1245: \label{g1dif}
1246: \eeq
1247: with initial condition
1248: \beq
1249: G\un(u,1)=\beta\,\frac{F(u)}{u}-2F'(u).
1250: \label{initg1}
1251: \eeq
1252: 
1253: Equation~(\ref{g1dif}) can again be solved by the method of Appendix B.
1254: We thus obtain
1255: \beq
1256: \frac{\dpar C(t,s)}{\dpar s}\approx s^{-(\beta-1)/2}\,\Phi\un(x),
1257: \label{dcsca}
1258: \eeq
1259: where
1260: \beq
1261: \Phi\un(x)=\frac{x^{-(\beta-1)/2}}{\zeta(\beta)}\int_0^\infty
1262: u^{1-\beta}\,G\un(u,x)\,\d u.
1263: \label{intphi1}
1264: \eeq
1265: This expression will not be made more explicit,
1266: as equation~(\ref{csca}) yields more directly
1267: \beq
1268: \Phi\un(x)=\frac{3-\beta}{2}\Phi(x)-x\,\frac{\d\Phi}{\d x},
1269: \label{phiphi1}
1270: \eeq
1271: i.e.,
1272: \beq
1273: \Phi\un(x)=x^{-\beta/2}\sum_{n=0}^\infty A\un_n\,x^{-n},
1274: \label{phi1exp}
1275: \eeq
1276: with
1277: \beq
1278: A\un_n=\left(n+\frac32\right)A_n.
1279: \eeq
1280: 
1281: The scaling behavior of the response function $R(t,s)$
1282: can be determined by the same approach.
1283: Looking for a two-variable scaling solution to equations~(\ref{dh}) in the
1284: scaling
1285: regime, of the form
1286: \be
1287: \h_k(t,s)\approx f_{k,\eq}\,t^{-1/2}\,G\de(u,x),
1288: \ee
1289: equations~(\ref{dh}) yield the partial differential equation
1290: \beq
1291: \left(x\frac{\dpar}{\dpar x}+\D-\frac12\right)G\de(u,x)=0,
1292: \label{g2dif}
1293: \eeq
1294: with initial condition
1295: \beq
1296: G\de(u,1)=\beta^2\frac{F(u)}{u}-\beta F'(u).
1297: \label{initg2}
1298: \eeq
1299: 
1300: We thus obtain
1301: \beq
1302: R(t,s)\approx s^{-(\beta-1)/2}\,\Phi\de(x),
1303: \label{rsca}
1304: \eeq
1305: where
1306: \beq
1307: \Phi\de(x)=\frac{x^{-(\beta-1)/2}}{\zeta(\beta)}\int_0^\infty
1308: u^{1-\beta}\,G\de(u,x)\,\d u,
1309: \label{intphi2}
1310: \eeq
1311: i.e., explicitly
1312: \beq
1313: \Phi\de(x)=x^{-\beta/2}\sum_{n=0}^\infty A\de_n\,x^{-n}.
1314: \label{phi2exp}
1315: \eeq
1316: The leading coefficient of this expansion reads
1317: \be
1318: A\de_0=\frac{\beta\,\pi^{1/2}\,2^{1-\beta}
1319: \Gamma\!\left(\frac{\beta+2}{2}\right)}
1320: {\Gamma\!\left(\frac{\beta+1}{2}\right)^2\zeta(\beta)}
1321: =\frac{3\beta(\beta+1)}{2(\beta+2)}\,A_0,
1322: \ee
1323: while the coefficient ratios are again rational functions of $\beta$:
1324: \be
1325: \frac{A\de_n}{A\de_0}
1326: =\frac{1}{(\beta+1)2^n}\prod_{j=0}^{n-1}(2j+\beta)\,\cdot\,
1327: \sum_{k=0}^n\frac{(-)^k(2k+\beta+1)}{(2k+1)k!(n-k)!}\prod_{\ell=0}^{k-1}
1328: \frac{2\ell+\beta+2}{2\ell+\beta+3},
1329: \ee
1330: i.e.,
1331: \be
1332: \frac{A\de_1}{A\de_0}=\frac{\beta(2\beta+1)}{6(\beta+1)},\qquad
1333: \frac{A\de_2}{A\de_0}=\frac{\beta(\beta+2)(8\beta^2+28\beta+9)}
1334: {120(\beta+1)(\beta+3)},\qquad\hbox{etc.}
1335: \ee
1336: 
1337: The scaling results~(\ref{dcsca}) and~(\ref{rsca})
1338: imply that the \fd ratio $X(t,s)$, defined in equation~(\ref{xdef}),
1339: only depends on the time ratio $x$ in the nonequilibrium scaling regime.
1340: We have indeed
1341: \beq
1342: X(t,s)\approx\X(x)=\frac{\Phi\de(x)}{\beta\Phi\un(x)}.
1343: \label{xsca}
1344: \eeq
1345: The scaling function $\X(x)$ turns out to be universal,
1346: whereas $\Phi(x)$, $\Phi\un(x)$, and $\Phi\de(x)$,
1347: which respectively enter equations~(\ref{csca}),~(\ref{dcsca}),~(\ref{rsca}),
1348: are only universal up to a scale fixing.
1349: (A definition of universal quantities
1350: in the present context will be recalled in the Discussion.)
1351: 
1352: Equations~(\ref{phi1exp}) and~(\ref{phi2exp}) yield the expression
1353: \beq
1354: \X(x)
1355: =\frad{\sum_{n=0}^\infty A\de_n\,x^{-n}}{\beta\sum_{n=0}^\infty A\un_n\,x^{-n}}
1356: =\sum_{n=0}^\infty\xi_n\,x^{-n},
1357: \label{xexp}
1358: \eeq
1359: which is again convergent for $x>1$.
1360: The limit \fd ratio, $X_\infty=\X(\infty)=\xi_0$, takes the simple value
1361: \beq
1362: X_\infty=\frac{\beta+1}{\beta+2}.
1363: \label{xinf}
1364: \eeq
1365: Equation~(\ref{xexp}) also yields
1366: \be
1367: \xi_1=\frac{\beta^2}{3(\beta+2)(\beta+3)},\qquad
1368: \xi_2=\frac{\beta^2(-2\beta^3+2\beta^2+78\beta+117)}
1369: {45(\beta+2)(\beta+3)^2(\beta+5)},\qquad\hbox{etc.}
1370: \ee
1371: 
1372: The behavior of the \fd ratio $\X(x)$
1373: close to equilibrium (i.e., for $x\to1$) is also of interest.
1374: Equations~(\ref{intphi1}) and~(\ref{intphi2}) for $x=1$,
1375: together with~(\ref{initg1}) and~(\ref{initg2}), imply
1376: \be
1377: \lim_{x\to1}\left(\beta\Phi\un(x)-\Phi\de(x)\right)
1378: =-\frac{\beta}{\zeta(\beta)}\,M_{F'}(\beta-2)
1379: =\frac{2^{1-\beta}\,\beta}{\Gamma\!\left(\frac{\beta+1}{2}\right)\zeta(\beta)}.
1380: \ee
1381: This expression, together with equations~(\ref{phieq}) and~(\ref{phiphi1}),
1382: yields
1383: \be
1384: \X(x)\approx1-\frac{2\,\Gamma\!\left(\frac{\beta}{2}\right)^2}
1385: {\pi\,\Gamma\!\left(\frac{\beta+1}{2}\right)
1386: \Gamma\!\left(\frac{\beta-1}{2}\right)}\,(x-1)^{(\beta-1)/2},
1387: \ee
1388: confirming that the \fd theorem is restored for $x\to1$,
1389: and explicitly giving the leading violation of that theorem
1390: in the scaling regime, which is proportional to $(\tau/s)^{(\beta-1)/2}$.
1391: 
1392: \section{Nonequilibrium dynamics in the condensed phase}
1393: 
1394: We finally turn to the long-time behavior of the two-time quantities
1395: $C(t,s)$, $R(t,s)$, and $X(t,s)$ in the condensed phase~($\rho>\rho_c$).
1396: 
1397: In Regime~II, we look for two-variable scaling solutions
1398: to equations~(\ref{dg}),~(\ref{dp}), and~(\ref{dh}),
1399: inspired by~(\ref{cofsca}), of the form
1400: \beq
1401: \g_k(t,s)\approx\frac{k}{t}\,G(u,x),\quad
1402: \p_k(t,s)\approx\frac{k}{t^2}\,G\un(u,x),\quad
1403: \h_k(t,s)\approx\frac{k}{t^2}\,G\de(u,x),
1404: \label{cosca}
1405: \eeq
1406: where $u=k\,t^{-1/2}$ and $x=t/s$.
1407: 
1408: The three scaling functions obey partial differential equations
1409: of the form~(\ref{gdif}),~(\ref{g1dif}), or~(\ref{g2dif}),
1410: where the differential operator $\D$ is replaced by
1411: \beq
1412: \D=-\frac{\d^2}{\d u^2}
1413: +\left(-\frac{u}{2}+A-\frac{\beta+2}{u}\right)\frac{\d}{\d u}
1414: -\frac{3}{2}+\frac{A}{u}.
1415: \eeq
1416: This differential operator is related to the Heun operator $\H$~(\ref{cohdef})
1417: and to the function $Y(u)$ defined in equation~(\ref{coeta}) by the conjugation
1418: \beq
1419: Y(u)\,\H=\D\,Y(u).
1420: \eeq
1421: Hence $\D$ and $\H$ share the same eigenvalues $E_n$ ($n=0,1,\dots$),
1422: which are not known explicitly, as already mentioned, except $E_0=0$.
1423: 
1424: The two-time scaling functions defined in~(\ref{cosca})
1425: are then determined in analogy with the critical case.
1426: We thus obtain
1427: \beq
1428: G(u,x)=Y(u)\sum_{n=0}^\infty a_n x^{-E_n-1/2}\,H_n(u),\qquad
1429: G\unde(u,x)=Y(u)\sum_{n=0}^\infty a\unde_n x^{-E_n+1/2}\,H_n(u),
1430: \label{coexpan}
1431: \eeq
1432: where the coefficients $a_n$, $a\unde_n$ are determined
1433: by the initial conditions
1434: \beq
1435: G(u,1)=F(u),\quad
1436: G\un(u,1)=\left(\frac{A}{u}-\frac{\beta}{u^2}\right)F(u)
1437: -\frac{2}{u}\,F'(u),\quad
1438: G\de(u,1)=-\frac{\beta}{u}\,F'(u).
1439: \label{cotwoi}
1440: \eeq
1441: 
1442: Using the expansions~(\ref{coexpan}),
1443: as well as the linear dependence of $F(u)$ in $(\rho-\rho_c)$
1444: (see equations~(\ref{coftoh}),~(\ref{coc})),
1445: we obtain the following scaling forms for the
1446: correlation and response functions:
1447: \beq
1448: \matrix{
1449: \ds{C(t,s)\approx s^{1/2}\,(\rho-\rho_c)\,\Phi(x),}\hfill&
1450: \ds{\Phi(x)=\sum_{n=0}^\infty A_n\,x^{-E_n},}\hfill\cr
1451: \ds{\frac{\dpar C(t,s)}{\dpar s}
1452: \approx s^{-1/2}\,(\rho-\rho_c)\,\Phi\un(x),}\qquad\hfill&
1453: \ds{\Phi\un(x)=\sum_{n=0}^\infty A\un_n\,x^{-E_n},}\hfill\cr
1454: \ds{R(t,s)\approx s^{-1/2}\,(\rho-\rho_c)\,\Phi\de(x),}\hfill&
1455: \ds{\Phi\de(x)=\sum_{n=0}^\infty A\de_n\,x^{-E_n}.}\hfill\cr
1456: }
1457: \eeq
1458: The scaling functions $\Phi(x)$, $\Phi\unde(x)$ only depend on $\beta$.
1459: They take finite limit values, both at $x=0$ and at $x=\infty$.
1460: The \fd ratio
1461: \be
1462: X(t,s)\approx\X(x)=\frac{\Phi\de(x)}{\beta\Phi\un(x)}
1463: \ee
1464: again only depends on the time ratio $x$
1465: in the nonequilibrium scaling regime.
1466: The whole scaling function $\X(x)$ is universal,
1467: just as in the critical case.
1468: In particular, the limit \fd ratio $X_\infty=\X(\infty)=A\de_0/A\un_0$
1469: in the condensed phase is universal, and only depends on $\beta$.
1470: Using equations~(\ref{coftoh}) and~(\ref{cotwoi}),
1471: we can derive the following expression for $X_\infty$
1472: in terms of the ground-state eigenfunction $H_0(u)$:
1473: \beq
1474: X_\infty=\frad
1475: {\int_0^\infty\left(\frac{u}{2}-A+\frac{\beta}{u}\right) H_0^2(u)\,\d u}
1476: {\int_0^{\infty^\vb}u\,H_0^2(u)\,\d u}.
1477: \label{coxinf}
1478: \eeq
1479: 
1480: The above expressions cannot be made more explicit in general.
1481: More quantitative results can be derived
1482: at high temperature and at low temperature, using results of section~3.2.
1483: 
1484: Let us consider the limit \fd ratio $X_\infty$, as given by
1485: equation~(\ref{coxinf}).
1486: In the high-temperature case, by~(\ref{cohih}) and~(\ref{cohia}), we obtain
1487: \beq
1488: X_\infty=\frac{1}{2}-(\pi-2)\beta+\cdots\qquad(\beta\to 0).
1489: \label{coxhi}
1490: \eeq
1491: In the low-temperature case, using~(\ref{coloset}) and~(\ref{coloh}),
1492: we obtain
1493: \beq
1494: X_\infty=\beta^{-1/2}-\frac{\beta^{-3/4}}{4}+\cdots\qquad(\beta\to\infty).
1495: \label{coxlo}
1496: \eeq
1497: 
1498: It is instructive to compare the limit \fd ratios
1499: corresponding to the critical point and to the condensed phase.
1500: The value of $X_\infty$ at criticality is given by the analytical
1501: expression~(\ref{xinf}).
1502: The value of $X_\infty$ in the condensed phase is obtained by
1503: a numerical evaluation of~(\ref{coxinf}):
1504: the ground-state eigenfunction $H_0(u)$ is obtained by numerically
1505: solving the differential equation~(\ref{cofdif})
1506: or~(\ref{cohdif}), and then used to evaluate
1507: the integrals entering the result~(\ref{coxinf}).
1508: The values thus obtained smoothly interpolate
1509: between the limiting laws~(\ref{coxhi}) and~(\ref{coxlo}).
1510: In Figure~\ref{fx},
1511: thick full lines show the physical values of the \fd ratios,
1512: while thin dashed lines show their continuation to high temperatures.
1513: Both \fd ratios start from $1/2$ at infinite temperature,
1514: and converge to the limit values $X_\infty=0$ and $X_\infty=1$
1515: at zero temperature, respectively in the condensed phase and at criticality
1516: (see Discussion).
1517: 
1518: \begin{figure}[htb]
1519: \begin{center}
1520: \includegraphics[angle=90,width=.8\linewidth]{fx.eps}
1521: \caption{\small
1522: Plot of the limit \fd ratio $X_\infty$ against inverse temperature $\beta$.
1523: Upper curve: critical point ($\beta>3$, $\rho=\rho_c$)
1524: (see~(\ref{xinf})).
1525: Lower curve: condensed phase ($\beta>2$, $\rho>\rho_c$)
1526: (see~(\ref{coxinf})).
1527: Thin dashed lines: continuation of the results to high temperature.}
1528: \label{fx}
1529: \end{center}
1530: \end{figure}
1531: 
1532: \section{Discussion}
1533: 
1534: At the onset of this paper we gave a comparative presentation
1535: of two classes of dynamical urn models,
1536: the Ehrenfest class and the Monkey class.
1537: All these models have simple static properties,
1538: because their Hamiltonian is a sum of contributions of independent boxes.
1539: They possess however interesting nonequilibrium dynamical properties,
1540: even in the mean-field geometry.
1541: The backgammon model~\cite{ritort,fr,gbm,gl}
1542: is a prototypical example of the Ehrenfest class, while
1543: model B of Reference~\cite{gbm}, and the zeta urn model~\cite{bia,dgc}
1544: which is the subject of the present work, are examples of the other class.
1545: 
1546: Let us come back to the role of statistics in the definition and
1547: properties of such models.
1548: The essential difference between the Ehrenfest class and the Monkey class
1549: indeed resides in matters related to a priori statistics.
1550: Statistics enters the dynamical definition of models:
1551: the proposed moves for the Ehrenfest class (respectively, the Monkey class)
1552: are chosen according to ball-box statistics (respectively, box-box statistics).
1553: Consistently, statistics also enters their equilibrium definition:
1554: the partition functions~(\ref{z2n}),~(\ref{zinfty})
1555: of the original Ehrenfest model, and of its generalization to $M$ urns,
1556: involve a factorial of the total number of particles in their denominators.
1557: Inverse factorials $1/N_i!$,
1558: taking into account equivalent labelings of particles within each box,
1559: are also involved in the evaluation of the partition function~(\ref{jstar}).
1560: 
1561: In statistical mechanics with Maxwell-Boltzmann statistics,
1562: the presence of inverse factorials has its origin in
1563: the indiscernibility of identical classical objects.
1564: These factorials are absent for the Monkey class,
1565: in which the populations $N_i$ are involved in flat sums,
1566: such as~(\ref{Zbb}), just as occupation numbers are in quantum-mechanical
1567: statistical mechanics with Bose-Einstein statistics.
1568: There is of course nothing quantum-mechanical in the urn models considered
1569: here.
1570: It can be said for short, following Reference~\cite{kim},
1571: that the equilibrium statistics of the Ehrenfest class is Maxwell-Boltzmann,
1572: while that of the Monkey class is Bose-Einstein.
1573: Table~1 illustrates this discussion.
1574: 
1575: \begin{table}[htb]
1576: \begin{center}
1577: \begin{tabular}{|c|c|c|}
1578: \hline
1579: Class&{\bf Ehrenfest}&{\bf Monkey}\\
1580: \hline
1581: Dynamical rule&ball-box&box-box\\
1582: \hline
1583: Equilibrium weight of box
1584: $i$&$\frad{p_{N_i}^\vb}{N_{i_{\vb_\vb}}!}$&$p_{N_i}$\\
1585: \hline
1586: Statistical-mechanical analogue&classical&quantum-mechanical\\
1587: &(Maxwell-Boltzmann)&(Bose-Einstein)\\
1588: \hline
1589: \end{tabular}
1590: \caption{\small
1591: Comparison of a priori statistics for the Ehrenfest class and the Monkey class
1592: of dynamical urn models: dynamical and equilibrium aspects.}
1593: \end{center}
1594: \end{table}
1595: 
1596: The main focus of this work concerns the nonequilibrium properties
1597: of the zeta urn model,
1598: with emphasis on the aging behavior of the two-time correlation
1599: and response functions of the fluctuating population of a given box,
1600: and of the corresponding \fd ratio.
1601: We considered successively the critical line ($\rho=\rho_c$) and the condensed
1602: phase ($\rho>\rho_c$).
1603: We summarize these two cases below.
1604: 
1605: The critical line in the temperature-density plane
1606: corresponds to a line of fixed points,
1607: parametrized by inverse temperature $\beta$~\cite{bia}.
1608: In other words, critical exponents, both static and dynamical,
1609: and more generally universal quantities, depend continuously on temperature.
1610: In the present context, a universal quantity ought to be independent:
1611: \begin{description}
1612: \item
1613: -- of the initial state (provided it is homogeneous and disordered,
1614: i.e., the $f_k(0)$ decay rapidly),
1615: \item
1616: -- of the specific form of the energy of each box
1617: (provided it diverges logarithmically, as $E(N_i)\approx\ln N_i$,
1618: for large occupation numbers),
1619: \item
1620: -- and of details of the dynamics (such as Metropolis versus heat-bath).
1621: \end{description}
1622: 
1623: Our results~(\ref{csca}) and~(\ref{rsca})
1624: for the two-time correlation and response function
1625: have the expected product form~(see~\cite{glcrit} and references therein),
1626: involving: a common non-universal prefactor,
1627: a negative power of the waiting time, related to the anomalous dimension
1628: of the observable, and a universal scaling function of the time ratio,
1629: or temporal aspect ratio, $x=t/s$.
1630: Accordingly, the \fd ratio~(\ref{xsca}), $X(t,s)\approx\X(x)$,
1631: only depends on $x$ in the nonequilibrium scaling regime.
1632: The limit value $X_\infty=\X(\infty)$ has been recently
1633: emphasized~\cite{glglau,glcrit} to be a new universal quantity,
1634: characteristic of nonequilibrium critical dynamics.
1635: For the zeta urn model, the result~(\ref{xinf})
1636: applies to the regular part of the critical line~($\beta>3$),
1637: so that $4/5<X_\infty<1$, as shown by the upper curve in Figure~\ref{fx}.
1638: This range is unusual for a critical system.
1639: Indeed, statistical-mechanical models such as ferromagnets
1640: are observed to have $0<X_\infty\le1/2$ at their critical point.
1641: The upper bound $X_\infty=1/2$, corresponding to the mean-field
1642: situation~\cite{glcrit},
1643: is also observed in a range of simpler models~\cite{ck,aging,glglau}.
1644: It is worth noticing that the backgammon model
1645: also has a high \fd ratio at its zero-temperature critical point,
1646: namely $X(t,s)\approx 1-C/(\ln s)^2$ for $s\ll t$,
1647: where the amplitude $C$ depends both on the observable
1648: and on the dynamical rule~\cite{fr,gl}.
1649: 
1650: The present analysis of condensation dynamics
1651: for $\rho>\rho_c$ extends and completes that begun in~\cite{dgc}.
1652: We have investigated various quantities related to the occupation probabilities
1653: in the scaling regime, describing the growing condensate.
1654: An approximate analysis of the Heun operator~(\ref{cohdef})
1655: has allowed us to obtain asymptotic expressions
1656: for various quantities at high and low temperature.
1657: The \fd ratio admits a non-trivial limit value $X_\infty$
1658: throughout the condensed phase, shown by the lower curve in Figure~\ref{fx}.
1659: Expression~(\ref{coxlo}) shows that $X_\infty\approx\beta^{-1/2}$
1660: slowly goes to zero at low temperature.
1661: This behavior is very different from that of conventional models.
1662: It is indeed currently accepted~\cite{xzero} that coarsening systems,
1663: such as ferromagnets quenched from a high-temperature initial state,
1664: have identically $X_\infty=0$ throughout their low-temperature phase,
1665: i.e., for any temperature below $T_c$.
1666: These unusual features of the zeta urn model are less of a surprise
1667: if one remembers that the condensation dynamics of the present model
1668: is basically different from a domain-growth or coarsening dynamics.
1669: In the latter case, phase separation takes place in a statistically homogeneous way, at least for an infinite system.
1670: To the contrary, in the present situation,
1671: condensation takes place in a very inhomogeneous fashion.
1672: The form of equation~(\ref{cofsca}) indeed demonstrates that
1673: the condensate of particles is shared
1674: by an ever decreasing fraction of boxes, scaling as $t^{-1/2}$,
1675: each of them having a population growing as $t^{1/2}$
1676: (until size effects eventually become important for $t\sim M^2$).
1677: 
1678: \newpage
1679: \appendix
1680: \section{Solving equation~(\ref{fdif}) by Mellin transformation}
1681: 
1682: The Mellin transformation provides an efficient alternative way of solving
1683: the differential equation~(\ref{fdif}).
1684: 
1685: We define the Mellin transform $M_f(z)$ of a function $f(u)$ by
1686: \beq
1687: M_f(z)=\int_0^\infty u^{-z-1}\,f(u)\,\d u,\qquad
1688: f(u)=\int\frac{\d z}{2\pi\i}\,u^z\,M_f(z).
1689: \label{mdef}
1690: \eeq
1691: 
1692: Equation~(\ref{fdif}), together with the boundary condition $F(0)=1$,
1693: is equivalent to the following functional equation
1694: \be
1695: \frac{M_{1-F}(z+2)}{M_{1-F}(z)}=\frac{z}{2(z+2)(\beta-1-z)}
1696: \ee
1697: for the Mellin transform of $1-F(u)$.
1698: The solution of this functional equation,
1699: with boundary condition $F(\infty)=0$, reads
1700: \beq
1701: M_{1-F}(z)=\frac{\Gamma\!\left(\frac{\beta+1-z}{2}\right)}
1702: {z\,2^z\,\Gamma\!\left(\frac{\beta+1}{2}\right)}\qquad(0<\re z<\beta+1).
1703: \label{melumf}
1704: \eeq
1705: Similarly, for $\re z<0$, the Mellin transform of $F(u)$,
1706: \beq
1707: M_F(z)=-\frac{\Gamma\!\left(\frac{\beta+1-z}{2}\right)}
1708: {z\,2^z\,\Gamma\!\left(\frac{\beta+1}{2}\right)}\qquad(\re z<0),
1709: \label{melf}
1710: \eeq
1711: is the formal opposite of equation~(\ref{melumf}).
1712: 
1713: We also give for further reference the expression of the Mellin transform
1714: of the derivative $F'(u)$:
1715: \beq
1716: M_{F'}(z)=-\underbrace{(z+1)M_{1-F}(z+1)}_{-1<\re z<\beta}
1717: =\underbrace{(z+1)M_F(z+1)}_{\re z<-1}
1718: =-\frac{\Gamma\!\left(\frac{\beta-z}{2}\right)}
1719: {2^{z+1}\,\Gamma\!\left(\frac{\beta+1}{2}\right)}
1720: \qquad(\re z<\beta).
1721: \label{melfp}
1722: \eeq
1723: 
1724: \section{Solving equation~(\ref{gdif}) by separation of variables
1725: and spectral superposition}
1726: 
1727: Partial differential equations such as~(\ref{gdif}),
1728: with given initial and boundary conditions,
1729: can be solved explicitly by the method of separation of variables
1730: and spectral superposition.
1731: 
1732: To do so, we need a basis of eigenfunctions of the differential operator
1733: $\D$ of equation~(\ref{ddef}).
1734: Inspired by the explicit form of the solution~(\ref{fex}), we set
1735: \be
1736: G(u)=u^{\beta+1}\,\e^{-u^2/4}\,L(v),\qquad v=u^2/4.
1737: \ee
1738: The eigenvalue equation $(\D-E)G(u)=0$ becomes
1739: \be
1740: v\frac{\d^2 L}{\d v^2}
1741: +\left(\frac{\beta+3}{2}-v\right)\frac{\d L}{\d v}+(E-1)L=0,
1742: \ee
1743: which is the differential equation
1744: obeyed by the Laguerre polynomials $L_n^\alpha(v)$~\cite{bateman},
1745: with $\alpha=(\beta+1)/2$ and $n=E-1$.
1746: 
1747: The eigenvalues of the operator $\D$ therefore read $E_n=n+1$,
1748: with $n=0,1,\dots$
1749: The associated eigenfunctions,
1750: \be
1751: G_n(u)=u^{\beta+1}\,\e^{-u^2/4}\,\L_n(u^2/4),
1752: \ee
1753: with~\cite{bateman}
1754: \beq
1755: \L_n(v)=\sum_{k=0}^n\frac{\Gamma\!\left(n+\frac{\beta+3}{2}\right)}
1756: {\Gamma\!\left(k+\frac{\beta+3}{2}\right)}\frac{(-v)^k}{k!(n-k)!},
1757: \label{lag}
1758: \eeq
1759: obey the orthogonality property
1760: \beq
1761: \int_0^\infty
1762: G_m(u)\,G_n(u)\,u^{-\beta}\,\e^{u^2/4}\,\d u=N_n\,\delta_{m,n},
1763: \qquad N_n=\frac{2^{\beta+2}\,\Gamma\!\left(n+\frac{\beta+3}{2}\right)}{n!}.
1764: \label{orthal}
1765: \eeq
1766: 
1767: The Mellin transform $M_{G_n}(z)$
1768: of these eigenfunctions can again be evaluated in closed form:
1769: \beq
1770: M_{G_n}(z)=\frac
1771: {2^{\beta-z}\,\Gamma\!\left(n+\frac{z+2}{2}\right)\,
1772: \Gamma\!\left(\frac{\beta+1-z}{2}\right)}
1773: {n!\,\Gamma\!\left(\frac{z+2}{2}\right)}
1774: \qquad(\re z<\beta+1),
1775: \label{melgn}
1776: \eeq
1777: where the normalization has been fixed by the condition~\cite{bateman}
1778: $\L_n(0)=1$.
1779: 
1780: The method of spectral superposition consists in looking
1781: for the solution $G(u,x)$ of equation~(\ref{gdif})
1782: as a linear superposition of the form
1783: \beq
1784: G(u,x)=\sum_{n=0}^\infty a_n x^{-(n+3/2)} G_n(u).
1785: \label{gexpand}
1786: \eeq
1787: The coefficients $a_n$ are determined by the initial condition~(\ref{initg}).
1788: Using the orthogonality property~(\ref{orthal}), we obtain
1789: \be
1790: a_n=\frac{1}{N_n}\int_0^\infty u^2\,F(u)\,\L_n(u^2/4)\,\d u.
1791: \ee
1792: Then, using~(\ref{lag}) and~(\ref{melgn}), we are left with
1793: \beq
1794: a_n=\frac{2^{1-\beta}}{\Gamma\!\left(\frac{\beta+1}{2}\right)}
1795: \,n!\sum_{k=0}^n\frac{(-)^k}{(2k+3)k!(n-k)!}
1796: \frac{\Gamma\!\left(k+\frac{\beta+4}{2}\right)}
1797: {\Gamma\!\left(k+\frac{\beta+3}{2}\right)}.
1798: \label{an}
1799: \eeq
1800: 
1801: \newpage
1802: \begin{thebibliography}{99}
1803: 
1804: \bibitem{ehr} P. and T. Ehrenfest, Phys. Zeit. {\bf 8} (1907), 311.
1805: 
1806: \bibitem{ks} F. Kohlrausch and E. Schr\"odinger, Phys. Zeit. {\bf 27}
1807: (1926), 306.
1808: 
1809: \bibitem{kac} M. Kac, Amer. Math. Monthly {\bf 54} (1947), 369.
1810: 
1811: \bibitem{kac2} M. Kac, in {\it Probability and Related Topics in Physical
1812: Sciences}, Lectures in Applied Mathematics, vol.~{\bf 1~A} (American
1813: Mathematical Society, 1959).
1814: 
1815: \bibitem{sieg} A.J.F. Siegert, Phys. Rev. {\bf 76} (1949), 1708.
1816: 
1817: \bibitem{hess} F.G. Hess, Amer. Math. Monthly {\bf 61} (1954), 323.
1818: 
1819: \bibitem{ritort} F. Ritort, Phys. Rev. Lett. {\bf 75} (1995), 1190.
1820: 
1821: \bibitem{fr} S. Franz and F. Ritort, Europhys. Lett. {\bf 31} (1995), 507;
1822: J. Stat. Phys. {\bf 85} (1996), 131; J. Phys. A {\bf 30} (1997), L359.
1823: 
1824: \bibitem{gbm} C. Godr\`eche, J.P. Bouchaud, and M. M\'ezard, J. Phys. A
1825: {\bf 28} (1995), L603.
1826: 
1827: \bibitem{gl} C. Godr\`eche and J.M. Luck, J. Phys. A {\bf 29} (1996), 1915;
1828: J. Phys. A {\bf 30} (1997), 6245; J. Phys. A {\bf 32} (1999), 6033.
1829: 
1830: \bibitem{bia} P. Bialas, Z. Burda, and D. Johnston, Nucl. Phys. B {\bf 493}
1831: (1997), 505; Nucl. Phys. B {\bf 542} (1999), 413; P. Bialas, L. Bogacz,
1832: Z. Burda, and D. Johnston, Nucl. Phys. B {\bf 575} (2000), 599.
1833: 
1834: \bibitem{dgc} J.M. Drouffe, C. Godr\`eche, and F. Camia, J. Phys. A {\bf 31}
1835: (1998), L19.
1836: 
1837: \bibitem{ck} L.F. Cugliandolo and J. Kurchan, J. Phys. A {\bf 27} (1994), 5749.
1838: 
1839: \bibitem{aging} For reviews, see: E. Vincent, J. Hammann, M. Ocio, J.P.
1840: Bouchaud, and L.F. Cugliandolo, in {\it Complex Behavior of Glassy Systems},
1841: Springer Lecture Notes in Physics {\bf 492} (1997), 184 (cond-mat/9607224);
1842: J.P. Bouchaud, L.F. Cugliandolo, J. Kurchan, and M. M\'ezard,
1843: in {\it Spin Glasses and Random Fields},
1844: Directions in Condensed Matter Physics, vol.~{\bf 12},
1845: ed. A.P. Young (World Scientific, Singapore, 1998) (cond-mat/9702070).
1846: 
1847: \bibitem{bray} A.J. Bray, Adv. Phys. {\bf 43} (1994), 357.
1848: 
1849: \bibitem{glglau} C. Godr\`eche and J.M. Luck, J. Phys. A {\bf 33} (2000), 1151.
1850: 
1851: \bibitem{glcrit} C. Godr\`eche and J.M. Luck, J. Phys. A {\bf 33} (2000), 9141.
1852: 
1853: \bibitem{heun} A. Ronveaux, {\it Heun's Differential Equations}
1854: (Oxford University Press, Oxford, 1995).
1855: 
1856: \bibitem{kim} B.J. Kim, G.S. Jeon, and M.Y. Choi, Phys. Rev. Lett. {\bf 76}
1857: (1996), 4648.
1858: 
1859: \bibitem{xzero} L.F. Cugliandolo, J. Kurchan, and L. Peliti, Phys. Rev. E
1860: {\bf 55} (1997), 3898; A. Barrat, Phys. Rev. E {\bf 57} (1998), 3629;
1861: L. Berthier, J.L. Barrat, and J. Kurchan, Eur. Phys. J. B {\bf 11} (1999), 635.
1862: 
1863: \bibitem{bateman} A. Erd\'elyi (ed.), {\it Higher Transcendental Functions}
1864: (The Bateman Manuscript Project) (McGraw-Hill, New-York, 1953).
1865: 
1866: \end{thebibliography}
1867: \end{document}
1868: