cond-mat0208034/urn.tex
1: \documentclass[aps,twocolumn,groupedaddress,showkeys,showpacs]{revtex4}
2: \usepackage[dvips]{graphicx}
3: \bibliographystyle{apsrev}
4: 
5: \begin{document}
6: 
7: \title{Analytic study of the urn model for separation of sand}
8: 
9: \author{G.\ M.\ Shim}
10: \author{B.\ Y.\ Park}
11: \author{Hoyun Lee}
12: \affiliation{Department of Physics, Chungnam National University, 
13: Daejeon 305-764, R.\ O.\ Korea}
14: 
15: \date{\today}
16: 
17: \begin{abstract}
18: We present an analytic study of the urn model for separation of sand 
19: recently introduced by Lipowski and Droz (Phys.\ Rev.\ E {\bf 65}, 
20: 031307 (2002)). 
21: We solve analytically the master equation and the first-passage problem. 
22: The analytic results confirm the numerical results obtained by 
23: Lipowski and Droz.
24: We find that the stationary probability distribution and the shortest 
25: one among the characteristic times are governed by the same {\it 
26: free energy}. We also analytically derive the form of the critical 
27: probability distribution on the critical line, 
28: which supports their results obtained by numerically calculating 
29: Binder cumulants (cond-mat/0201472).
30: 
31: \end{abstract}
32: 
33: \pacs{45.70.--n,68.35.Rh}
34: %pacs:granular systems, phase transitions and critical phenomena
35: \keywords{granular, urn model, master equation, first-passage problem, critical phenomena, symmetry breaking.}
36: 
37: \maketitle
38: 
39: %%%%%%%%%%%%%%%%%%%%%%%% section 1 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40: \section{INTRODUCTION}
41: 
42: A granular system exhibits extremely rich phenomena, which has recently
43: attracted extensive studies. One of such interesting phenomena is the 
44: spatial separation of shaken sand \cite{Schlichting96}. Sand in a box 
45: separated into two equal parts by a wall that allows the transfer of 
46: sand through its narrow slit prefers to aggregate more in one side 
47: under certain conditions. 
48: 
49: Eggers explained the emergence of symmetry breaking using a hydrodynamic 
50: approach \cite{Eggers99}. The key idea is to introduce the effective 
51: temperature taking into account the inelastic collisions for granular 
52: material.
53: 
54: Lipowski and Droz proposed a dynamic model to explain the essence of
55: the phenomena \cite{Lipowski02a}. The model is a certain generalization 
56: of the Ehrenfest model \cite{Ehrenfest90}. Interestingly this model 
57: shows a spontaneous symmetry breaking in contrast to other 
58: generalizations of the Ehrenfest model. 
59: They derived the master equation and found in a numerical way 
60: the phase diagram that 
61: displays a rich structure like continuous and discontinuous transitions 
62: as well as a tricritical point. They also numerically solved the 
63: first-passage problem to find exponential or algebraic divergences.
64: 
65: Thanks to its simplicity, the model allows analytic approaches.
66: In this paper, we present the results of this analytic study to the 
67: master equation and the first-passage problem addressed by Lipowski 
68: and Droz. These not only confirm their numerical results but 
69: also give us some insights in the nature of the discontinuous transition 
70: in the stationary probability distribution.
71: We also analytically derive the form of the critical probability 
72: distribution on the critical line.
73: 
74: The paper is organized as follows. 
75: In Sec.~\ref{sec:II} we briefly review the model and its master equation. 
76: In Sec.~\ref{sec:III} we present the analytic solution of the master 
77: equation in the thermodynamic limit and the analytic expression of 
78: the stationary probability distribution. The form of the stationary
79: probability distribution on the critical line is also derived.
80: In Sec.~\ref{sec:IV} we analytically solve the first-passage problem.
81: Detailed analysis on the behavior of the characteristic times is given 
82: in Sec.~\ref{sec:V}. Section \ref{sec:VI} is devoted to conclusions 
83: and discussions.
84: 
85: %%%%%%%%%%%%%%%%%%%%%%%% sec 2 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
86: \section{MODEL AND ITS MASTER EQUATION} \label{sec:II}
87: 
88: The model introduced by Lipowski and Droz \cite{Lipowski02a} is 
89: defined as follows. $N$ particles are distributed between two urns, 
90: and the number of particles in each urn is denoted as $M$ and $N-M$, 
91: respectively. 
92: At each time of updates one of the $N$ particles is randomly chosen. 
93: Let $n$ be a fraction of the total number of particles in the urn 
94: that the selected particles belongs to. 
95: With probability $\exp(-\frac{1}{T(n)})$ the selected particle moves 
96: to the other urn. $T(n)$ represents the effective temperature of an
97: urn with particles $nN$ that measures the thermal fluctuations 
98: of the urn.
99: Lipowski and Droz chose the temperature as $T(n) = T_0+\Delta(1-n)$. 
100: 
101: It is easy to derive the master equation for the probability
102: distribution $p(M,t)$ that there are $M$ particles in a given urn at 
103: time $t$ \cite{Lipowski02a}
104: \begin{eqnarray} 
105:    p(M,t+1) &=& F\bigl(\frac{N-M+1}{N}\bigr) p(M-1,t)
106:   \nonumber\\
107:             &&  + F\bigl(\frac{M+1}{N}\bigr) p(M+1,t)
108:   \nonumber\\
109:             &&  + \Bigl[  1-F\bigl(\frac{M}{N}\bigr)
110:                     -F\bigl(\frac{N-M}{N}\bigr) \Bigr] p(M,t)
111: \,,\label{eq:master}\end{eqnarray}
112: where $F(n) = n \exp(-\frac{1}{T(n)})$ measures the flux of 
113: particles leaving the given urn.
114: Here we introduced for convenience the notations $p(-1,t)=p(N+1,t)=0$.
115: 
116: The difference in the occupancy of the urns can be represented by 
117: the particle excess $\epsilon = \frac{M}{N}-\frac12$. The time 
118: evolution of the averaged particle excess
119: $e(t)=\langle\epsilon\rangle_t=\sum_M (\frac{M}{N}-\frac12) p(M,t)$
120: is governed by 
121: \begin{equation}\label{eq:averagedexcess}
122:    e(t+1) = e(t)+\frac{1}{N}\bigl\langle {\cal F}(\epsilon) 
123:                            \bigr\rangle_t
124: \,,\end{equation}
125: where ${\cal F}(\epsilon)= F(\frac12-\epsilon)-F(\frac12+\epsilon)$
126: measures the net flux of particles in the given urn. 
127: One conventionally takes the unit of time in such a way that there
128: is a single update per a particle on average. Therefore we scale
129: the time by $N$. Expanding Eq.~(\ref{eq:averagedexcess}) with respect 
130: to $\frac{1}{N}$, and using the mean--field approximation in evaluating
131: the average, we get
132: \begin{equation}\label{eq:differentialexcess}
133:   \frac{d}{dt}e(t) = {\cal F}\bigl(e(t)\bigr)
134: \,.\end{equation}
135: Note that the stationary solution of Eq.~(\ref{eq:differentialexcess}) 
136: is determined by zero points of ${\cal F}(\cdot)$. 
137: The stable stationary solutions are given by zero points of ${\cal F}(\cdot)$
138: with a negative slope, which we call as the stable fixed points while the
139: unstable ones are given by those with a positive slope, which we call as
140: the unstable fixed points.
141: 
142: Detailed analysis on the existence of the stable stationary solutions of
143: Eq.~(\ref{eq:differentialexcess}) was done by Lipowski and Droz
144: \cite{Lipowski02a}. We here display their phase diagram in
145: Fig.~\ref{fig:phasediagram} to make our paper
146: as self-contained as possible. 
147: The stable symmetric solution ($\epsilon=0$) exists in region I, III, and
148: IV while the stable asymmetric solution ($\epsilon>0$) exists in region II,
149: III and IV.
150: 
151: 
152: \begin{figure}[tbp]%--------------------------------------------%
153:    \includegraphics[width=7cm]{phasediagram.eps} 
154:    \caption{\label{fig:phasediagram} Phase diagram of the urn model
155:             \cite{Lipowski02a}.
156:             The symmetric solution vanishes continuously on the solid line
157:             while the asymmetric one disappears discontinuously on the 
158:             dotted line. The transition of the behavior of the stationary
159:             probability distribution is denoted by the dashed line.}
160: \end{figure}%--------------------------------------------------%
161: 
162: 
163: %%%%%%%%%%%%%%%%%%%%%%%% section 3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
164: \section{THE SOLUTION OF THE MASTER EQUATION} \label{sec:III}
165: 
166: We are mainly interested in investigating the properties of the 
167: infinite system.
168: Consider the thermodynamic limit $N \rightarrow \infty$ with
169: $\frac{M}{N}=\frac12+\epsilon$ being fixed. 
170: Representing the probability distribution by $\epsilon$ instead
171: of $M$, and expanding Eq.~(\ref{eq:master}) with respect to 
172: $\frac{1}{N}$, and keeping the terms up to the first order,
173: we arrive at the expression
174: \begin{eqnarray}
175:    p(\epsilon,t+1) &=& p(\epsilon,t)+\frac{1}{N}\Bigl[
176:                         \bigl(  F^\prime(\frac12+\epsilon)
177:                                +F^\prime(\frac12-\epsilon)
178:                         \bigr) p(\epsilon,t)
179:    \nonumber\\
180:                    &&   +\bigr(  F(\frac12+\epsilon)
181:                                -F(\frac12-\epsilon)
182:                         \bigr) \frac{\partial}{\partial \epsilon}
183:                                p(\epsilon,t)
184:                     \Bigr]  
185: \,.\label{eq:masterexpand}\end{eqnarray}
186: 
187: Scaling again the time by $N$, expanding Eq.~(\ref{eq:masterexpand}) 
188: with respect to $\frac{1}{N}$, and noting that the second term in the 
189: right-handed side can be combined into a total derivative with respect to
190: $\epsilon$, we finally obtain the partial differential equation
191: \begin{equation}\label{eq:continuousmaster}
192:     \frac{\partial}{\partial t}p(\epsilon,t)
193:    + \frac{\partial}{\partial \epsilon}\bigl[ 
194:          {\cal F}(\epsilon)p(\epsilon,t)
195:      \bigr] = 0 
196: \,.\end{equation}
197: 
198: Note that ${\cal F}(\cdot)$ is zero at a finite number of points.
199: The solution of Eq.~(\ref{eq:continuousmaster}) can be found in the 
200: intervals that do not include those points. At each interval, it would 
201: be convenient to introduce a new variable
202: \begin{equation}\label{eq:parametrization}
203:     \lambda(\epsilon) = \int_{\epsilon_0}^\epsilon \frac{dx}{{\cal F}(x)}
204: \,,\end{equation}
205: where $\epsilon_0$ is a certain point in the interval. 
206: Figure \ref{fig:map} shows a typical behavior of the mapping. 
207: We also displayed the map for $|\epsilon|>\frac12$ by analytic
208: continuation. This is necessary since the solution of 
209: Eq.~(\ref{eq:continuousmaster}) is of wave nature.
210: (See Eq.~(\ref{eq:wavesolution}) below.)
211: $\lambda$ increases as
212: $\epsilon$ approaches to the stable fixed points while it decreases as 
213: $\epsilon$ approaches to the unstable fixed points.
214: \begin{figure}[tbp]%--------------------------------------------%
215:    \includegraphics[width=7cm]{map.eps} 
216:    \caption{\label{fig:map} $\lambda(\epsilon)$ for $\Delta=0.3, T_0=0.2$. 
217:             The mapping for $\epsilon>\frac12$ is analytically continuated.}
218: \end{figure}%--------------------------------------------------%
219: With the help of this parameterization and setting 
220: $R(\lambda,t)={\cal F}(\epsilon)p(\epsilon,t)$, 
221: Eq.~(\ref{eq:continuousmaster}) now takes the form
222: \begin{equation}\label{eq:wave}
223:     \frac{\partial}{\partial t}R(\lambda,t)
224:    +\frac{\partial}{\partial \lambda}R(\lambda,t) = 0
225: \,.\end{equation}
226: Note that Eq.~(\ref{eq:wave}) is in fact a half part of
227: the wave equations so that its solution is written as 
228: $R(\lambda,t) = f(\lambda-t)$ with $f(\cdot)$ being an arbitrary
229: differentiable function. It represents a wave that moves to the direction
230: of increasing $\lambda$ as time evolves, which means that the system
231: moves to stable fixed points. 
232: The solution of the original partial differential equation
233: (\ref{eq:continuousmaster}) now reads
234: \begin{equation}\label{eq:wavesolution}
235:     p(\epsilon,t) = \frac{ f\bigl(\lambda(\epsilon)-t\bigr) }{
236:                            {\cal F}(\epsilon) } 
237: \,.\end{equation}
238: 
239: From the initial probability distribution $p_0(\epsilon) = p(\epsilon,0)$
240: we can determine the function $f(\cdot)$. So we get 
241: \begin{equation}\label{eq:solutiondistrib} 
242:    p(\epsilon,t) = \frac{{\cal F}(\epsilon_t)}{{\cal F}(\epsilon)}
243:                    p_0(\epsilon_t)
244: \,,\end{equation}
245: where $\epsilon_t$ is given by the relation
246: \begin{equation}\label{eq:inversemap}
247:     \lambda(\epsilon_t) = \lambda(\epsilon)-t
248: \,.\end{equation}
249: Here it should be understood that $\epsilon_t$ is to be chosen in
250: the same interval where $\epsilon$ belongs to. 
251: Furthermore $p_0(\epsilon)=0$ for $|\epsilon|>\frac12$ is assumed since
252: $\epsilon_t$ in Eq.~(\ref{eq:solutiondistrib}) can be larger 
253: (or smaller) than $\frac12$ (-$\frac12$). This happens because of the
254: nature of the wave solution.
255: Using the mapping from $\epsilon$ to $\epsilon_t$, it is straightforward to
256: show that the total probability is conserved:
257: \begin{equation}
258:        \int_{-\frac12}^\frac12 p(\epsilon,t)d\epsilon
259:      = \int_{-\frac12}^\frac12 p_0(\epsilon_t)d\epsilon_t
260:      = 1
261: \,.\end{equation} 
262: 
263: The shape of the probability distribution is distorted by a ratio 
264: ${\cal F}(\epsilon_t)/{\cal F}(\epsilon)$ so that it accumulates
265: at the nearest stable fixed points.
266: In Eq.~(\ref{eq:solutiondistrib}), the ratio approaches zero as time 
267: evolves unless $\epsilon$ is on a stable fixed point. 
268: As a consequence, in the long time limit $t \rightarrow \infty$ 
269: the probability distribution becomes a sum of delta peaks at the stable
270: fixed points  $\epsilon_i$
271: \begin{equation}\label{eq:probdistinfty}
272:     p(\epsilon,\infty) = \sum_i p_i \delta(\epsilon-\epsilon_i)
273: \,,\end{equation}
274: where $p_i$ are the sum of the initial probabilities 
275: in two intervals adjacent to its point $\epsilon_i$.
276: We would like to point out that the system is not ergodic and its dynamical
277: phase space is decomposed into disconnected sectors. Each sector is
278: associated with a stable fixed point and is separated by the unstable fixed
279: points.
280: 
281: The fixed point condition ${\cal F}(\epsilon)=0$ is equivalent to Eq.~(4)
282: in Ref.~\cite{Lipowski02a}, where Lipowski and Droz analyzed in detail the
283: condition and their results are summarized in the phase diagram (See
284: Fig.~\ref{fig:phasediagram}.). 
285: We would like to mention that 
286: in regions III and IV in Fig.~\ref{fig:phasediagram} both the
287: symmetric and the asymmetric solutions exist together. In fact, either
288: solution can be realized by choosing an appropriate initial configuration.
289: Lipowski and Droz distinguished regions III and IV according to the 
290: different behaviors of the stationary probability distribution
291: in their numerical process of taking the limit $N \rightarrow \infty$. 
292: In region III there appear two delta peaks for the
293: asymmetric solutions while in region IV there appears only the central delta 
294: peak for the symmetric solution. It is contradictory to our result 
295: Eq.~(\ref{eq:probdistinfty}) where any
296: delta peaks for the stable fixed points can appear depending on the initial
297: configurations.
298: 
299: To resolve this contradiction and understand the nature of the transition
300: between regions III and IV, we take another limit in the master
301: equation (\ref{eq:master}), namely take the long time limit $t \rightarrow
302: \infty$ before we take the limit $N \rightarrow \infty$. This limit may not
303: properly reflect the properties of the infinite system. 
304: Since the infinite system is not ergodic as we showed above,
305: changing the order of taking limits $N \rightarrow \infty$ and
306: $t \rightarrow \infty$ may not yield the same result. 
307: Anyway it seems that in their simulations about the stationary probability 
308: distribution, Lipowski and Droz took the limit $t \rightarrow \infty$ 
309: for a finite system size $N$, and then extrapolating the results 
310: to $N \rightarrow \infty$.
311: 
312: Let's first take the long time limit of $t \rightarrow \infty$ in 
313: Eq.~(\ref{eq:master}). In this limit we may drop off the time dependence in
314: the probability distribution, which now takes the form
315: \begin{eqnarray}
316:    p(N) &=& F\bigl(\frac1N\bigr)p(N-1)+\bigl(1-F(1)\bigr)p(N) 
317:  \nonumber\\
318:    p(M) &=& F\bigl(\frac{N-M+1}{N}\bigr)p(M-1)
319:  \nonumber\\
320:         &&    +F\bigl(\frac{M+1}{N}\bigr)p(M+1)
321:  \nonumber\\
322:         &&    +\Bigl[ 1-F\bigl(\frac{M}{N}\bigr)-
323:                       F\bigl(\frac{N-M}{N}\bigr)\Bigr]p(M)
324:  \nonumber\\
325:          && \mbox{for $M=N-1,\ldots,2,1$}
326:  \nonumber\\
327:    p(0) &=& F\bigl(\frac1N\bigr)p(1)+\bigl(1-F(1)\bigr)p(0)
328: \,.\label{eq:stationaryprob}\end{eqnarray}
329: The first equation in Eq.~(\ref{eq:stationaryprob}) allows us to rewrite 
330: $p(N)$ in terms of $p(N-1)$, which in turn allows to rewrite $p(N-1)$ 
331: in terms of $p(N-2)$, and so on. Therefore we find
332: \begin{eqnarray}
333:     p(M) &=& \frac{ F\bigl(\frac{N-M+1}{N}\bigr) }{
334:                     F\bigl(\frac{M}{N}\bigr)} p(M-1)
335:   \nonumber\\
336:          &=& p(0) \prod_{i=1}^M \frac{ F\bigl(\frac{N-i+1}{N}\bigr) }{
337:                    F\bigl(\frac{i}{N}\bigr)}
338: \label{eq:stationaryrecursive}\end{eqnarray}
339: for $M=N,\ldots,2,1$. $p(0)$ appears as an overall factor to normalize
340: the probabilities so that we get
341: \begin{equation}
342:    p(0) = \Bigl[ 1+\sum_{M=1}^N \prod_{i=1}^M 
343:             \frac{F\bigl(\frac{N-i+1}{N}\bigr)}{F\bigl(\frac{i}{N}\bigr)}
344:           \Bigr]^{-1}
345: \,.\end{equation}
346: Now let's take the limit $N \rightarrow \infty$.
347: With $\frac{M}{N} = \frac12+\epsilon$, and 
348: $\frac{i}{N}= \frac12+x$, and scaling the probability distribution by
349: $N$, the stationary probability distribution for large $N$ now becomes
350: \begin{equation}\label{eq:statcontinprob}
351:    p_s(\epsilon) \approx \frac{ e^{NG(\epsilon)} }{
352:                  \int_{-\frac12}^\frac12 dx \, e^{NG(x)} }
353: \,,\end{equation}
354: where 
355: \begin{equation}
356:     G(\epsilon) = \int_{-\frac12}^\epsilon dx \,
357:                   \bigl[ \log F(\frac12-x) - \log F(\frac12+x) \bigr]
358: \,.\end{equation}
359: 
360: In the limit $N \rightarrow \infty$, the main contribution to the
361: stationary probability distribution comes only from the maximum of
362: $G(\cdot)$, and it becomes delta peaks. 
363: The maximum of $G(\epsilon)$ occurs when
364: \begin{equation}
365:     G^\prime(\epsilon) =  \log F(\frac12-\epsilon) 
366:                          -\log F(\frac12+\epsilon) 
367:                        = 0
368: \,.\end{equation}
369: Since $G^\prime(\epsilon)$ is the difference of logarithms of 
370: $ F(\frac12-\epsilon) $ and $ F(\frac12+\epsilon) $, 
371: both $G(\cdot)$ and ${\cal F}(\cdot)$ share the similar qualitative 
372: properties. For example, the maximum of both functions occurs at 
373: the same stable fixed points. 
374: Note that in region II only the two asymmetric solutions 
375: with positive and negative particle excesses are stable,  and have the
376: same maximum  while in region I only the symmetric solution is stable. 
377: Therefore the stationary probability distribution has the double peaks 
378: in region II and  only the central peak in region I. 
379: In region III and IV both the
380: symmetric and the asymmetric solutions are stable so that the maximum
381: of $G(\epsilon)$ should be determined by comparing its values at the
382: stable fixed points. The crossover of the maximum point occurs when
383: both values coincide. This implies that the transition between
384: the double peaks and the central single peak in the probability
385: distribution is determined by the condition
386: $\Delta G = G(\epsilon_a)-G(0) = 0$, where $\epsilon_a$ is the nonzero
387: stable fixed point. This condition yields a line that separate two
388: regions III and IV.
389: 
390: It is very interesting to see that $-G(\cdot)$ resembles the free
391: energy of the equilibrium systems, and the transition between two
392: {\it phases} is determined by the condition that the free energies of
393: both phases are equal. Furthermore a certain characteristic time
394: behaves differently in two phases, as Lipowski and Droz numerically
395: found \cite{Lipowski02a}. We will show an analytic relation between
396: them in next section.
397: 
398:  
399: 
400: %%%%%%%%%%%%%%%%%%%%%%%% section 4 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
401: \section{CHARACTERISTIC TIMES}\label{sec:IV}
402: 
403: Lipowski and Droz defined an averaged first-passage time $\tau(M)$
404: needed for a configuration with $M$ particles in an urn (and $N-M$
405: particles in the other urn) to reach the symmetric configuration
406: ($M=\frac{N}{2}$) \cite{Lipowski02a}. 
407: They obtained the relations among the averaged characteristic 
408: times from the dynamical rules as
409: \begin{eqnarray}
410:    && \hspace{-2em}\tau(M) = F\bigl(\frac{M}{N}\bigr)\bigl[\tau(M-1)+1\bigr]
411:    \nonumber\\
412:    && \hspace{1em}  +F\bigl(\frac{N-M}{N}\bigr)\bigl[\tau(M+1)+1\bigr]
413:    \nonumber\\
414:    && \hspace{1em} +\Bigl[1-F\bigl(\frac{M}{N}\bigr)
415:                     -F\bigl(\frac{N-M}{N}\bigr)\Bigr]
416:                 \bigl[\tau(M)+1\bigr]
417: \label{eq:chartimerelation}\end{eqnarray}
418: for $M=N, N-1, \ldots, \frac{N}{2}+1$. Here it is understood that the
419: term associated with $\tau(N+1)$ does not appear. (In fact, its
420: coefficient $F(0)$ vanishes.) By definition of the characteristic
421: times, $\tau(\frac{N}{2})=0$ and $\tau(N-M) = \tau(M)$.
422: 
423: Defining the difference of successive characteristic times as 
424: $\Delta \tau(M)=\tau(M)-\tau(M-1)$, Eq.~(\ref{eq:chartimerelation}) can
425: be rewritten as
426: \begin{equation}
427:    \Delta \tau(M) = \frac{1}{F\bigl(\frac{M}{N}\bigr)}\Bigl[
428:                     1+F\bigl(\frac{N-M}{N}\bigr)\Delta\tau(M+1) \Bigr]
429: \,.\end{equation}
430: By applying this relation repeatedly until
431: $\Delta\tau(N)=\frac{1}{F(1)}$ is reached, we get the expression
432: \begin{equation}\label{eq:Dtaurelation}
433:    \Delta\tau(M) = \frac{1}{F\bigl(\frac{M}{N}\bigr)}\Bigl[
434:                    1+\sum_{i=1}^{N-M} \prod_{j=1}^i 
435:                     \frac{ F\bigl(\frac{N-M-j+1}{N}\bigr) }{
436:                             F\bigl(\frac{M+j}{N}\bigr) }
437:                   \Bigr]
438: \,.\end{equation}
439: Since $\tau(\frac{N}{2})=0$ by definition, we immediately get
440: $\tau(\frac{N}{2}+1) = \Delta\tau(\frac{N}{2}+1)$, which is given
441: by Eq.~(\ref{eq:Dtaurelation}) with $M=\frac{N}{2}+1$. 
442: By successively adding $\Delta\tau(M)$, we get the general expression 
443: for $\tau(M)$ for $M=N,N-1,\ldots,\frac{N}{2}+1$:
444: \begin{equation}
445:    \tau(M) = \sum_{k=\frac{N}{2}+1}^M \frac{1}{F\bigl(\frac{k}{N}\bigr)}
446:                  \Bigl[1+\sum_{i=1}^{N-k} \prod_{j=1}^i 
447:                     \frac{ F\bigl(\frac{N-M-j+1}{N}\bigr) }{ 
448:                             F\bigl(\frac{M+j}{N}\bigr) }
449:                   \Bigr]
450: \,.\end{equation}
451:                   
452: We are mainly interested in the behavior of $\tau(M)$ as $N$ increases.
453: Again we scale the characteristic times by $N$. 
454: With $\frac{M}{N}=\frac12+\epsilon$ being fixed, and introducing
455: $\frac{k}{N}=\frac12+x, \frac{i}{N}=y-x, \frac{j}{N}=z-x$, 
456: the summations for large $N$ can be replaced by integrations so that
457: the characteristic times for $\epsilon>0$ takes the form
458: \begin{equation}
459:    \tau(\epsilon) \approx N \int_0^\epsilon dx \int_x^\frac12 dy \,
460:                           e^{NH(x,y)}
461: \end{equation}
462: with $H(x,y) = G(y)-G(x)$. 
463: The longest characteristic time $\tau(N)$ corresponds to
464: $\tau(\epsilon=\frac12)$.  The shortest one $\tau(\frac{N}{2}+1)$ 
465: corresponding to $\tau(\epsilon=0$) is, in general, smaller than 
466: $\tau(\epsilon)$ with positive $\epsilon$ 
467: by an order of magnitude. It is necessary to deal with it separately.
468: We get
469: \begin{equation}
470:    \tau(\frac{N}{2}+1) \approx  \int_0^\frac12 dy \,
471:                           e^{NH(0,y)}
472: \,.\end{equation}
473: 
474: Since $H(0,y)=G(y)-G(0)$, both the shortest characteristic time
475: $\tau(\frac{N}{2}+1)$ and the stationary probability distribution 
476: $p_s(\epsilon)$ have essentially the same functional dependence on
477: $G(\cdot)$. Therefore it is not surprising that behaviors of both 
478: quantities for large $N$ are closely related. 
479: However it is not clear why they are.
480: 
481: %%%%%%%%%%%%%%%%%%%%%%%% section 5 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
482: \section{ANALYSIS OF THE CHARACTERISTIC TIMES}\label{sec:V}
483: 
484: We first consider the behavior of $\tau(\frac{N}{2}+1)$, the shortest
485: one among the characteristic times. For large $N$, the main
486: contribution to $\tau(\frac{N}{2}+1)$ comes from the maximum point
487: $y_m$ of $H(0,y)=G(y)-G(0)$, or $G(y)$, 
488: which was dealt in Sec.~\ref{sec:III}. 
489: 
490: In region I and IV the maximum occurs at $y_m=0$ corresponding to the
491: symmetric configuration, while in region II and III it occurs at 
492: $y_m>0$ corresponding to the asymmetric configurations. 
493: Therefore the maximum is zero in region I and IV while it is positive 
494: in region II and III.
495: 
496: When $y_m=0$, we may expand $H(0,y)$ around $y=0$ to get
497: \begin{eqnarray}
498:     H(0,y) &\approx& -2\Bigl[1-\frac{\Delta/2}{(T_0+\Delta/2)^2}\Bigr] y^2
499:   \nonumber\\
500:            && -\frac43\Bigl[1-\frac{3(\Delta/2)^2}{(T_0+\Delta/2)^4}
501:                       \Bigr] y^4 + \ldots
502: \,.\label{eq:Hyexpand0}\end{eqnarray}
503: Therefore it yields $\tau(\frac{N}{2}+1) \sim N^{-\frac12}$
504: as long as the coefficient of the first term in
505: Eq.~(\ref{eq:Hyexpand0}) is negative. This is the case in region I and
506: IV. The coefficient vanishes when 
507: $T_0 = \sqrt{\frac{\Delta}{2}}-\frac{\Delta}{2}$, 
508: which corresponds to the critical line found by Lipowski and Droz
509: \cite{Lipowski02a}. On this line, we get 
510: \begin{equation}\label{eq:scalingbehavior}
511:    H(0,y) \approx  -\frac43\bigl(1-\frac32\Delta\bigr) y^4
512:                    -\frac{32}{15}\bigl(1-\frac54\Delta^2\bigr)y^6
513:                    +\ldots
514: \,,\end{equation}
515: which yields $\tau(\frac{N}{2}+1) \sim N^{-\frac14}$ for
516: $\Delta < \frac23$, and $\tau(\frac{N}{2}+1) \sim N^{-\frac16}$ at
517: $\Delta = \frac23$ corresponding to the tricritical point.
518: 
519: When $y_m>0$, the maximum is positive so that 
520: $\tau(\frac{N}{2}+1) \sim N^{-\frac12}e^{\alpha N}$ (with $\alpha$ being
521: a positive constant), that is, it diverges exponentially.
522: 
523: Now let's investigate the behavior of the longest characteristic time
524: $\tau(N)$ or $\tau(\epsilon=1)$. Again the main contribution comes
525: from the maximum point of $H(x,y)$ in region restricted by three lines
526: $y=x, x=0, y=\frac12$. Note that $y \ge x$ in the region.
527: Interestingly the maximum point $(x_m,y_m)$ is closely related with 
528: the fixed points of ${\cal F}(\epsilon)$. 
529: 
530: In region I, $\epsilon=0$ is only the fixed point so that $x_m=y_m=0$, 
531: and the maximum is zero. Expanding $H(x,y)$ about this point, we get
532: \begin{equation}\label{eq:HxyExpand0}
533:   H(x,y) \approx -2\Bigl[1-\frac{\Delta/2}{(T_0+\Delta/2)^2}\Bigr]
534:                    (y^2-x^2)+\ldots
535: \,.\end{equation}
536: The coefficient of the leading term in Eq.~(\ref{eq:HxyExpand0}) is
537: negative only if $T_0 > \sqrt{\frac{\Delta}{2}}-\frac{\Delta}{2}$, that
538: is, above the critical line. Changing the variables $x, y$ to the polar
539: coordinates $r, \theta$ and scaling the radial coordinate $r$ by
540: $\sqrt{N}$, we arrive at 
541: \begin{eqnarray}
542:   \tau(N) &\approx& \int_0^{c\sqrt{N}} dr \, r
543:                   \int_{\frac{\pi}{4}}^{\frac{\pi}{2}} d\theta \, 
544:                   \exp\Bigl[-2\Bigl(1-\frac{\Delta/2}{(T_0+\Delta/2)^2}
545:                               \Bigr) 
546:                \nonumber \\
547:           && \hspace{9em}\times\, r^2\cos(2\theta)\Bigr]
548: \,.\end{eqnarray}
549: Here $c$ is a constant that gives the upper bound of the integration 
550: over $r$. The contribution from the neighborhood of $\theta=\frac{\pi}{4}$
551: yields a logarithmic divergence. Fig.~\ref{fig:tauNlogN} shows a typical 
552: behavior of characteristic time $\tau(N)$ as a function of $N$ 
553: for several values of $\Delta$ with $T_0=0.2$. The first two uppermost lines
554: stand for $\tau(N)$ in region II, and the others represent that in region
555: I. We conclude that in region I $\tau(N)$ diverges logarithmically 
556: as $N$ increases.
557: \begin{figure}[tbp]%--------------------------------------------%
558:    \includegraphics[width=7cm]{tauNlogN.eps} 
559:    \caption{\label{fig:tauNlogN} characteristic time  $\tau(N)$ as a 
560:             function of $N$ for $T_0=0.2$ and $\Delta=1.72, 1.744067, 
561:             1.77,1.8,2.0$ (from the top).}
562: \end{figure}%--------------------------------------------------%
563: 
564: As we see in Eq.~(\ref{eq:HxyExpand0}), the leading term vanishes on
565: the critical line. On this line we need to expand more. So
566: \begin{eqnarray}
567:    H(x,y) &\approx&  -\frac43\bigl(1-\frac32\Delta\bigr) (y^4-x^4)
568:        \nonumber\\
569:           &&         -\frac{32}{15}\bigl(1-\frac54\Delta^2\bigr)(y^6-x^6)
570:                       +\ldots
571: \,.\end{eqnarray}
572: Again changing the variables $x, y$ to the polar coordinates and scaling
573: the radial coordinate appropriately (by $N^\frac14$ or $N^\frac16$),
574: we conclude that $\tau(N)$ diverges algebraically as $N^\frac12$ on
575: the critical line and as $N^\frac23$ at the tricritical point.
576: 
577: In regions II, III, and IV 
578: there appear many fixed points among which we can always find one 
579: with $y_m>x_m$ and $G(y_m)>G(x_m)$. There, the maximum
580: $H(x_m,y_m)$ is positive, and $\tau(N)$ diverges exponentially
581: as $N$ increases. We would like to point out that the situation is
582: different from that of $\tau(\frac{N}{2}+1)$.
583: In fact, it corresponds to the case with $x_m$ being fixed to zero.
584: 
585: Finally let's consider the behavior of $\tau(N)$ on the line
586: separating two regions I and IV. As we approach this line from region
587: IV, the nonzero fixed points merge to disappear at
588: $\epsilon=\epsilon_1>0$. That is, we can find a positive $\epsilon_1$
589: such that ${\cal F}(\epsilon_1) = {\cal F}^\prime(\epsilon_1) = 0$.
590: On this line the maximum point is given by $x_m=y_m=\epsilon_1$ so that 
591: the maximum $H(x_m,y_m)$ is zero. 
592: Expanding $H(x,y)$ around this point, we get
593: \begin{equation}
594:      H(x,y) \approx -\frac16 G^{\prime\prime\prime}(\epsilon_1)
595:                     \bigl((y-\epsilon_1)^3-(x-\epsilon_1)^3\bigr)
596:                     +\ldots
597: \,.\end{equation}
598: (Here we don't write down $ G^{\prime\prime\prime}(\epsilon_1)$
599: explicitly since it is not important as far as it is positive.) 
600: Note that the leading order is the third instead of the fourth 
601: as in case of the critical line. 
602: The reason is that $G(\epsilon)$ is not symmetric about 
603: $\epsilon = \epsilon_1$ while it is symmetric about $\epsilon = 0$.
604: Consequently $\tau(N)$ on this line diverges algebraically as
605: $N^\frac13$.
606: 
607: 
608: 
609: %%%%%%%%%%%%%%%%%%%%%%%% section 6 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610: \section{CONCLUSIONS}\label{sec:VI}
611: 
612: We analytically investigate the urn model introduced by Lipowski and
613: Droz \cite{Lipowski02a}. We exactly solve the master equation of the
614: model in the thermodynamic limit and find how the probability
615: distribution evolves. 
616: In the long time limit, the probability distribution becomes 
617: delta peaks only at the stable fixed points. 
618: In fact the ergodicity of the dynamics is broken so that the dynamical
619: phase space is decomposed into disconnected sectors separated by the
620: unstable fixed points. The strength of a delta peak is equal to the
621: sum of initial probabilities in the disconnected sector it belongs to.
622: 
623: We also solve exactly the stationary probability distribution where we
624: take the long time limit before we take thermodynamic limit.
625: Regardless of the initial probability distribution it shows double
626: peaks or a single central peak depending on the parameters of the
627: system. The final formula of the stationary probability distribution 
628: resembles that of the equilibrium systems, where the transition from 
629: the doubles peaks to the single peak is determined by the condition 
630: that {\it free energies} of two phases become equal.
631: 
632: Recently Lipowski and Droz \cite{Lipowski02b}
633: numerically calculated  Binder cumulants
634: of the urn model to find that the critical probability distribution
635: has the form $p(x) \sim e^{-x^4}$ on the critical line, and
636: $p(x) \sim e^{-x^6}$ on the tricritical point, where $x$ is the rescaled
637: order parameter proportional to the particle excess $\epsilon$.
638: As we showed, $G(\epsilon) \sim H(0,\epsilon)$ and
639: its behavior on the critical line (including the tricritical point)
640: is given by Eq.~(\ref{eq:scalingbehavior}). Therefore the critical
641: probability distribution has the form $p(\epsilon) \sim \exp[
642: -\frac43(1-\frac32\Delta)\epsilon^4]$ on the critical line,
643: and $p(\epsilon) \sim \exp[-\frac{128}{135}\epsilon^6]$ on the
644: tricritical point.
645: Our analytic result supports their numerical result.
646: 
647: The first-passage problem is analytically solved. 
648: Interestingly both the shortest characteristic time 
649: and the stationary probability distribution are governed by the
650: same {\it free energy}. 
651: Therefore the behavior of the shortest characteristic
652: time and the properties of the stationary probability distribution are
653: closely related.
654: The analytic results on the behavior of the characteristic times 
655: support the numerical results of Lipowski and Droz \cite{Lipowski02a}.
656: 
657: Finally it would be very interesting to understand why both the stationary 
658: probability distribution and the shortest characteristic time are 
659: governed by the same {\it free energy}.
660: It would also be of interest to extend our analytic study to many-urn 
661: models and other types of urn models.
662: 
663: 
664: 
665: 
666: %%%%%%%%%%%%%%%%%%%%%%%% references %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
667: \begin{thebibliography}{99}
668: \bibitem{Schlichting96} H.\ J.\ Schlichting and V.\ Nordmeier,
669:                         Math.\ Naturwiss.\ Unterr.\ {\bf 49}, 323 (1996).
670: \bibitem{Eggers99} J.\ Eggers,
671:                    Phys.\ Rev.\ Lett.\ {\bf 83}, 5322 (1999).
672: \bibitem{Lipowski02a} A.\ Lipowski and M.\ Droz, 
673:                       Phys.\ Rev.\ E {\bf 65}, 031307 (2002).
674: \bibitem{Ehrenfest90} P.\ Ehrenfest and T.\ Ehrenfest,
675:                       {\it The Conceptual Foundations of the 
676:                       Statistical Approach in Mechanics} 
677:                       (Dover, New York, 1990);
678:                   M.\ Kac and J.\ Logan,
679:                   in {\it Fluctuation Phenomena}, 
680:                   edited by E.\ W.\ Montroll and J.\ L.\ Lebowitz 
681:                   (North-Holland, Amsterdam, 1987).
682: \bibitem{Lipowski02b} A.\ Lipowski and M.\ Droz, cond-mat/0201472.
683: \end{thebibliography}
684:              
685: \end{document}
686: