0705.0907/1.tex
1: \documentclass[12pt]{iopart}
2: \usepackage{graphicx}
3: \eqnobysec
4: 
5: \newcommand{\beq}{\begin{equation}}
6: \newcommand{\beqa}{\begin{eqnarray}}
7: \newcommand{\eeq}{\end{equation}}
8: \newcommand{\eeqa}{\end{eqnarray}}
9: \newcommand{\bfa}{{\bf a}}
10: \newcommand{\bfb}{{\bf b}}
11: \newcommand{\bfd}{{\bf d}}
12: \newcommand{\bfe}{{\bf e}}
13: \newcommand{\bfBB}{{\hskip -.3mm}{\bf B{\hskip -.6mm}B}{\hskip -.3mm}}
14: \newcommand{\bfSS}{{\hskip -.1mm}{\bf S{\hskip -.1mm}S}{\hskip -.1mm}}
15: \newcommand{\bg}{{\rm bg}}
16: \newcommand{\bigmean}[1]{\left\langle#1\right\rangle}
17: \renewcommand{\d}{{\rm d}}
18: \newcommand{\dr}{\delta\rho}
19: \newcommand{\ds}{\displaystyle}
20: \renewcommand{\e}{{\rm e}}
21: \newcommand{\frad}[2]{\displaystyle{\displaystyle#1\over\displaystyle#2}}
22: \newcommand{\gl}{{\rm gl}}
23: \newcommand{\hdr}{\widehat{\delta\rho}}
24: \renewcommand{\i}{{\rm i}}
25: \newcommand{\loc}{{\rm loc}}
26: \renewcommand{\max}{{\rm max}}
27: \newcommand{\mean}[1]{\langle#1\rangle}
28: \newcommand{\prob}{\mathop{\rm Prob}\nolimits}
29: \newcommand{\BB}{{\rm BB}}
30: \newcommand{\BS}{{\rm BS}}
31: \newcommand{\J}{{\bf J}}
32: \newcommand{\N}{{\cal N}}
33: \newcommand{\SB}{{\rm SB}}
34: \renewcommand{\SS}{{\rm SS}}
35: 
36: \begin{document}
37: 
38: \title{Structure of the stationary state of the asymmetric target process}
39: 
40: \author{J M Luck and C Godr\`eche}
41: 
42: \address{Service de Physique Th\'eorique\footnote{URA 2306 of CNRS},
43: CEA Saclay, 91191 Gif-sur-Yvette cedex, France}
44: 
45: \begin{abstract}
46: We introduce a novel migration process, the target process.
47: This process is dual to the zero-range process (ZRP) in the sense that,
48: while for the ZRP the rate of transfer of a particle
49: only depends on the occupation of the departure site,
50: it only depends on the occupation of the arrival site for the target process.
51: More precisely, duality associates to a given ZRP a unique target process,
52: and vice-versa.
53: If the dynamics is symmetric, i.e., in the absence of a bias,
54: both processes have the same stationary-state product measure.
55: In this work we focus our interest on the situation where
56: the latter measure exhibits a continuous condensation transition
57: at some finite critical density $\rho_c$, irrespective of the dimensionality.
58: The novelty comes from the case of asymmetric dynamics,
59: where the target process has a nontrivial fluctuating stationary state,
60: whose characteristics depend on the dimensionality.
61: In one dimension, the system remains homogeneous at any finite density.
62: An alternating scenario however prevails in the high-density regime:
63: typical configurations consist of long
64: alternating sequences of highly occupied and less occupied sites.
65: The local density of the latter is equal to~$\rho_c$
66: and their occupation distribution is critical.
67: In dimension two and above, the asymmetric target process exhibits a phase
68: transition at a threshold density $\rho_0$ much larger than~$\rho_c$.
69: The system is homogeneous at any density below $\rho_0$,
70: whereas for higher densities it exhibits an extended condensate
71: elongated along the direction of the mean current,
72: on top of a critical background with density $\rho_c$.
73: \end{abstract}
74: 
75: \pacs{05.40.-a, 02.50.Ey, 05.70.Ln}
76: 
77: \eads{\mailto{jean-marc.luck@cea.fr},
78: \mailto{claude.godreche@cea.fr}}
79: 
80: \maketitle
81: 
82: \section{Introduction}
83: 
84: In recent years many studies have been devoted to
85: nonequilibrium statistical-mechanical models yielding condensation,
86: such as zero-range processes
87: (ZRP)~\cite{loan,evans1,wis1,cg,gross,evans2,glcond},
88: dynamical urn models~\cite{zeta1,zeta2,zeta3,barc,lux},
89: and mass transport models~\cite{maj}.
90: In all these models the condensate manifests itself by
91: the macroscopic occupation of a single site
92: by a finite fraction of the whole available mass.
93: 
94: The ZRP is the simplest of these models.
95: It is a driven diffusive system with multiple occupations,
96: such that particles hop from site to site on a lattice,
97: with a rate which only depends on the occupation of the departure site.
98: The distribution of the particles among the sites
99: in the stationary state is given by a product measure,
100: which is explicitly known in terms of the rate defining the model,
101: irrespective of the geometry of the system
102: and of the asymmetry of the dynamics~\cite{evans2,spitz,andj}.
103: This property of the stationary-state measure of the ZRP
104: favors the condensation phenomenon.
105: The product structure indeed prevents the condensate
106: from being shared by more than one site~\cite{gross,glcond,maj}.
107: 
108: Dynamical urn models, also called migration processes in the probabilistic
109: li\-te\-ra\-ture~\cite{kelly}, can be viewed as generalizations of the ZRP,
110: where the rate at which a particle is transferred
111: from a departure site to an arrival site
112: now depends on the occupations of both sites.
113: Migration processes do not have a stationary-state product measure in general.
114: 
115: In the present work we introduce a special migration process,
116: the {\it target process}.
117: This process is novel to our knowledge.
118: It has no stationary-state product measure, except in the case
119: of a symmetric dynamics, i.e., in the absence of a bias,
120: leading to an equilibrium stationary state.
121: The class of target processes is dual to the class of ZRP,
122: in the sense that the roles of the departure and arrival sites are interchanged:
123: the rate basically depends on the occupation of the departure (source) site
124: for a ZRP,
125: and on the occupation of the arrival (target) site for a target process.
126: More precisely, to a given ZRP is associated by duality a unique target process,
127: and vice-versa.
128: In the case of a symmetric dynamics,
129: these two dual processes have the same stationary-state product measure.
130: 
131: Our aim is to study the structure of the nonequilibrium stationary state
132: of the asymmetric target process,
133: and especially the fate of the condensation phenomenon.
134: In Section~\ref{stp} we give a detailed definition
135: of the class of target processes.
136: We then focus our attention onto the particular target process
137: which is dual to the `canonical ZRP'
138: studied e.g.~in~\cite{evans1,wis1,cg,gross}.
139: In the case of symmetric dynamics, the target process thus constructed
140: has the same stationary-state product measure as the ZRP,
141: and therefore the same continuous condensation transition.
142: In the asymmetric case, however, the stationary-state measure of the target
143: process is not a product measure,
144: and exhibits non-trivial correlations in general.
145: This basic difference between the ZRP and the target process
146: manifests itself more drastically at high density and in low dimensionality.
147: Sections~\ref{1D1} and~\ref{1D2} are devoted to a thorough theoretical
148: and numerical study of the one-dimensional situation.
149: The system remains homogeneous at any finite density,
150: and presents an alternating structure which is more and more pronounced
151: as the density is increased.
152: The two-dimensional asymmetric target process
153: on the square lattice is the subject of Section~\ref{2D}.
154: It exhibits an unconventional type of condensation phenomenon,
155: with a transition at a threshold density $\rho_0$ much larger than~$\rho_c$,
156: and an extended condensate elongated along the direction of the bias.
157: An analogous scenario presumably generically holds on other lattices
158: and in higher dimension as well.
159: Section~\ref{quatre} contains a Discussion.
160: 
161: \section{The target process}
162: \label{stp}
163: 
164: \subsection{Migration processes and urn models: a reminder}
165: 
166: The definition of a migration process appeared first in the
167: probabilistic literature (see e.g.~\cite{kelly}).
168: For the time being, we restrict the discussion to the one-dimensional case.
169: Consider a system of $N$ particles distributed among~$M$ sites,
170: with periodic boundary conditions.
171: Let $N_m$ be the number of particles at site $m=1,\dots,M$.
172: A migration process (or dynamical urn model) is defined by the following
173: Markovian dynamics:
174: (i) a departure (source) site $d$ is chosen at random,
175: uniformly among the $M$ sites of the system;
176: (ii) an arrival (target) site $a$ is chosen among the neighbors of $d$.
177: To be specific, the right neighbor ($a=d+1$) is chosen with probability $p$,
178: whereas the left neighbor ($a=d-1$) is chosen with
179: the complementary probability $q=1-p$;
180: (iii) a particle is transferred from site $d$ to site $a$
181: at a rate $W_{k,l}$ which only depends on the occupations
182: $k=N_d$ and $l=N_a$ of the two sites involved.
183: Of course, one has
184: \beq
185: W_{0,l}=0,
186: \label{zcons}
187: \eeq
188: since no particle can be removed from an empty site.
189: 
190: A one-dimensional migration process
191: is therefore entirely defined by the bias $p$ and the rate $W_{k,l}$.
192: A natural question is the following:
193: {\it What are the conditions on the bias~$p$ and
194: the rate $W_{k,l}$ for a one-dimensional migration process
195: to have a stationary-state product measure?}
196: This question was first addressed in~\cite{cocozza}
197: (for a review, see~\cite{lux}).
198: 
199: In this context, a stationary-state product measure means
200: that the probability of any configuration of the system
201: in its stationary state has the form
202: \beq
203: P(N_1,\dots,N_M)=\frac{1}{Z_{M,N}}\;p_{N_1}\dots p_{N_M}
204: \;\delta(N_1+\cdots+N_M,N),
205: \label{pm}
206: \eeq
207: where the factors $p_k$ are arbitrary,
208: and the partition function $Z_{M,N}$ reads
209: \beq
210: Z_{M,N}=\sum_{N_1,\dots,N_M}p_{N_1}\dots p_{N_M}\;\delta(N_1+\cdots+N_M,N).
211: \eeq
212: 
213: The answer to the above question is as follows:
214: 
215: \begin{itemize}
216: \item
217: For symmetric dynamics, i.e., when $p=1/2$,
218: the stationary state has a product measure characterized by the factor $p_k$,
219: if the rate $W_{k,l}$ obeys the condition
220: \beq
221: p_{k+1}p_lW_{k+1,l}=p_kp_{l+1}W_{l+1,k}.
222: \label{m1}
223: \eeq
224: The resulting stationary state is an equilibrium state.
225: The relation~(\ref{m1}) expresses detailed balance
226: with respect to this equilibrium state.
227: 
228: \item
229: For asymmetric dynamics, i.e., when $p\neq1/2$,
230: the stationary-state has a product measure
231: if the rate $W_{k,l}$ obeys both~(\ref{m1}) and the following condition:
232: \beq
233: W_{k,l}-W_{k,0}=W_{l,k}-W_{l,0}.
234: \label{m2}
235: \eeq
236: The stationary state is a genuine nonequilibrium steady state.
237: The condition~(\ref{m1}) does not express detailed balance any longer,
238: albeit a weaker property~\cite{lux},
239: referred to as pairwise balance~\cite{pairw}.
240: \end{itemize}
241: 
242: The partition function $Z_{M,N}$ associated with the product measure~(\ref{pm})
243: can be rewritten, using an integral representation of
244: the Kronecker delta function, as
245: \beq
246: Z_{M,N}=\oint\frac{\d z}{2\pi\i z^{N+1}}\,P(z)^M,
247: \label{contour}
248: \eeq
249: where the generating series of the factors $p_k$ reads
250: \beq
251: P(z)=\sum_{k\ge0}p_kz^k.
252: \eeq
253: The product measure~(\ref{pm}) is therefore characterized by $M$, $N$,
254: and the factor $p_k$ or, equivalently, by the generating function $P(z)$.
255: For a homogeneous system in the thermodynamic limit,
256: where $M$ and $N$ are simultaneously large,
257: with a fixed density $\rho=N/M$ of particles per site,
258: the contour integral in~(\ref{contour}) can be evaluated
259: by the saddle-point method.
260: The saddle-point value $z$, which is to be identified with the fugacity
261: in the grand canonical ensemble,
262: is related to the density $\rho$ through the equation
263: \beq
264: \frac{z P'(z)}{P(z)}=\rho.
265: \label{col}
266: \eeq
267: The distribution $f_k=\prob\{N_1=k\}$
268: of the occupations of any given site of the system
269: can be derived by summing the probability~(\ref{pm}) over $N_2,\dots,N_M$.
270: We thus obtain
271: \beq
272: f_k=p_k\,\frac{Z_{M-1,N-k}}{Z_{M,N}}.
273: \eeq
274: In the thermodynamic limit, this expression simplifies to
275: \beq
276: f_k=\frac{p_kz^k}{P(z)}.
277: \label{fdef}
278: \eeq
279: 
280: \subsection{The example of the zero-range process}
281: 
282: In the present context, the ZRP appears as the special
283: case of a migration process where
284: the rate $W_{k,l}$ only depends on the occupation of the departure site:
285: \beq
286: W_{k,l}=u_k,
287: \label{zrpdef}
288: \eeq
289: with $u_0=0$, by virtue of~(\ref{zcons}).
290: The condition~(\ref{m2}) is then automatically satisfied,
291: irrespective of the bias $p$.
292: Equation~(\ref{m1}) yields the following relation between the rate~$u_k$
293: and the factor $p_k$:
294: \beq
295: p_k=u_{k+1}p_{k+1},
296: \label{urel}
297: \eeq
298: up to a multiplicative constant,
299: which we set equal to unity by an appropriate choice of time unit.
300: The corresponding factor $p_k$
301: can be expressed in terms of the rate $u_k$ as follows:
302: \beq
303: p_0=1,\qquad p_k=\frac{1}{u_1\dots u_k}\qquad(k\ge1).
304: \label{pzrp}
305: \eeq
306: Reciprocally, to a given stationary-state product measure
307: characterized by the factor~$p_k$, there corresponds a unique ZRP dynamics
308: (up to a choice of time unit), whose rate reads
309: \beq
310: u_k=\frac{p_{k-1}}{p_k}\qquad(k\ge1).
311: \label{udef}
312: \eeq
313: 
314: \subsection{Characterization of migration processes with stationary-state
315: product measure}
316: 
317: We now give an explicit characterization of the migration processes
318: which admit a stationary-state product measure.
319: 
320: For symmetric dynamics, and for a given factor $p_k$,
321: the most general form of the rate obeying~(\ref{m1}) reads
322: \beq
323: W_{k,l}=\frac{p_{k-1}}{p_k}\,S_{k-1,l}\qquad(k\ge1),
324: \label{wsym}
325: \eeq
326: where the ratio $p_{k-1}/p_k$ is nothing but the rate $u_k$
327: of the corresponding ZRP, given by~(\ref{udef}),
328: whereas~$S_{k,l}$ is a symmetric function of $k$ and~$l$:
329: \beq
330: S_{k,l}=S_{l,k}.
331: \eeq
332: Besides the factor~$p_k$, the rate $W_{k,l}$ depends on an arbitrary
333: symmetric function~$S_{k,l}$ of {\it two} indices.
334: 
335: For asymmetric dynamics, and for a given factor~$p_k$,
336: the most general solution of~(\ref{m1}) and~(\ref{m2}) is determined
337: by the one-dimensional array of rates $\alpha_k=W_{k,0}$~\cite{kls3}.
338: Note that $\alpha_k$ is the rate at which an empty site is refilled,
339: by receiving one particle from a non-empty neighboring site
340: containing $k\ge1$ particles, and that~(\ref{zcons}) implies $\alpha_0=0$.
341: The above property can be shown as follows.
342: Consider all the indices $k$ and $l$ for a fixed value of the sum $k+l=n$,
343: and introduce the quantities
344: \beq
345: A_k=p_kp_{n-k}W_{k,n-k}\qquad(k=0,\dots,n).
346: \eeq
347: Equations~(\ref{m1}) and~(\ref{m2}) respectively become
348: \beq
349: A_k=A_{n+1-k},\qquad A_k-p_kp_{n-k}\alpha_k=A_{n-k}-p_kp_{n-k}\alpha_{n-k}.
350: \eeq
351: Combining these two equations yields
352: \beq
353: A_{k+1}-A_k=p_kp_{n-k}(\alpha_{n-k}-\alpha_k).
354: \eeq
355: The solution of this inhomogeneous difference equation
356: with initial value $A_0=0$ reads
357: \beq
358: A_k=\sum_{m=1}^kp_{k-m}p_{n-k+m}(\alpha_{n-k+m}-\alpha_{k-m}).
359: \eeq
360: We are thus left with the following expression for the rate $W_{k,l}$:
361: \beq
362: W_{k,l}=\frac{1}{p_kp_l}\sum_{m=1}^kp_{k-m}p_{l+m}(\alpha_{l+m}-\alpha_{k-m}).
363: \label{wasym}
364: \eeq
365: Besides the factor $p_k$,
366: the rate $W_{k,l}$ depends on an arbitrary function
367: $\alpha_k$ of {\it one} index.
368: 
369: \subsection{Definition of the target process}
370: 
371: We define the target process as the migration process where the rate
372: \beq
373: W_{k,l}=(1-\delta_{k,0})v_l
374: \label{tpdef}
375: \eeq
376: essentially depends on the occupation $l$ of the arrival (target) site.
377: Note the dual character of this definition with respect to the definition of
378: the ZRP, in that the roles of the departure and arrival sites
379: have been interchanged.
380: There is, however, a key difference between the two models,
381: coming from the presence of the constraint~(\ref{zcons}).
382: The latter,
383: which explicitly enters~(\ref{tpdef}) through the factor $(1-\delta_{k,0})$,
384: implies that the rate of the target process actually also bears some
385: dependence on the occupation $k$ of the departure site,
386: as it is constrained to vanish if $k=0$.
387: For the ZRP the same constraint does not
388: change the fact that the rate~$u_k$ only depends on the departure site;
389: it just imposes $u_0=0$.
390: 
391: For symmetric dynamics,
392: the target process always has a stationary-state product measure,
393: for any choice of the rate $v_k$.
394: Equation~(\ref{m1}) yields the following relation between the rate $v_k$
395: and the factor $p_k$:
396: \beq
397: p_{k+1}=v_kp_k,
398: \label{vrel}
399: \eeq
400: up to a multiplicative constant,
401: which is again set to unity by an appropriate choice of time unit.
402: The factor $p_k$ of the stationary-state measure
403: can thus be expressed in terms of the rate $v_k$ as follows:
404: \beq
405: p_0=1,\qquad p_k=v_0\dots v_{k-1}\qquad(k\ge1).
406: \label{pvdef}
407: \eeq
408: Reciprocally, to a given stationary-state product measure,
409: characterized by the factor~$p_k$,
410: there corresponds a unique symmetric target process
411: (up to a choice of time unit), whose rate reads
412: \beq
413: v_k=\frac{p_{k+1}}{p_k}.
414: \label{vdef}
415: \eeq
416: 
417: Equations~(\ref{urel}) and~(\ref{vrel}) show that the ZRP with rate $u_k$
418: and the target process with rate $v_k$ have the same stationary-state
419: product measure, i.e., the same factor $p_k$, if the rates obey
420: \beq
421: v_k=\frac{1}{u_{k+1}},
422: \label{uvdual}
423: \eeq
424: up to a multiplicative constant.
425: A target process and a ZRP related by this condition
426: are hereafter named {\it dual} to each other.
427: 
428: It is interesting to consider a more general class of migration processes,
429: where the rate has the form
430: \beq
431: W_{k,l}=(1-\delta_{k,0})u_k\,v_l.
432: \label{wgen}
433: \eeq
434: For symmetric dynamics, this model again has a stationary-state product
435: measure for any choice of $u_k$ and $v_l$.
436: Equation~(\ref{m1}) indeed yields
437: \beq
438: u_{k+1}p_{k+1}=v_kp_k,
439: \label{genrel}
440: \eeq
441: up to a multiplicative constant.
442: The factor $p_k$
443: can thus be expressed in terms of the $u_k$ and $v_k$ as follows:
444: \beq
445: p_0=1,\qquad p_k=\frac{v_0\dots v_{k-1}}{u_1\dots u_k}\qquad(k\ge1).
446: \eeq
447: This expression shows that the factor $p_k$
448: only depends on the ratio $v_k/u_{k+1}$.
449: This class of processes interpolates between the ZRP and the target process,
450: which are respectively recovered as the special cases where $v_l=1$ and $u_k=1$.
451: We finally notice that the rate~(\ref{tpdef}) of the target process,
452: and more generally the rate~(\ref{wgen}), is of the form~(\ref{wsym}),~with
453: \beq
454: S_{k,l}=v_kv_l.
455: \eeq
456: 
457: For asymmetric dynamics, the condition~(\ref{m2}) is very stringent.
458: The target process has a stationary-state product measure
459: if and only if the rate $v_k$ only assumes two values,
460: according to whether $k$ is zero or~not:
461: \beq
462: v_k=\left\{\matrix{v_0\hfill&(k=0),\cr v\hfill&(k\ge1).}\right.%}
463: \label{two}
464: \eeq
465: This also holds for the more general process defined by the rate~(\ref{wgen}).
466: 
467: For a generic asymmetric target process,
468: where the rate $v_k$ is not of the form~(\ref{two}),
469: the stationary-state measure is not a product measure.
470: It is not known explicitly, and can be expected to be a non-trivial
471: correlated measure in general.
472: 
473: \subsection{The `canonical target process' considered in this work}
474: 
475: The condensation phenomenon in the ZRP is usually investigated
476: using the rate~\cite{evans1,wis1,cg,gross}:
477: \beq
478: u_k=1+\frac{b}{k}\qquad(k\ge1),
479: \label{ucan}
480: \eeq
481: where the control parameter $b$ is a measure of the strength of interactions.
482: The minimality and exemplarity of this choice of rate suggest to call this
483: model the `canonical ZRP for condensation', or `canonical ZRP', for short.
484: 
485: Throughout the following we focus our attention onto
486: the target process dual to the canonical ZRP.
487: Its rate $v_k$ is therefore related to the rate~(\ref{ucan})
488: by the duality relation~(\ref{uvdual}).
489: We thus obtain
490: \beq
491: v_k=\frac{k+1}{k+b+1}\qquad(k\ge0).
492: \label{vcan}
493: \eeq
494: We name this process the `canonical target process'.
495: 
496: For symmetric dynamics,
497: it has already been shown above that the target process has the same
498: stationary-state product measure as the ZRP.
499: It therefore exhibits the same condensation transition.
500: We now give a brief reminder of the properties of this stationary-state measure.
501: 
502: In the absence of interactions ($b=0$),
503: the rate reads $u_k=v_k=1$, irrespective of $k$.
504: We have therefore $p_k=1$, so that $P(z)=1/(1-z)$.
505: The fugacity $z$ and the density $\rho$ are related by
506: \beq
507: z=\frac{\rho}{\rho+1},\qquad\rho=\frac{z}{1-z}.
508: \eeq
509: The distribution of the occupations~(\ref{fdef}) is a geometric distribution:
510: \beq
511: f_k=(1-z)z^k.
512: \label{fexpo}
513: \eeq
514: 
515: In the situation where $b$ is positive,
516: the rate $u_k$ is a decreasing function of the occupation $k$,
517: so that particles hop less easily out of more occupied sites.
518: Accordingly, the rate $v_k$ is an increasing function of $k$,
519: hence particles hop preferentially towards more occupied sites.
520: The rate~(\ref{ucan}) or~(\ref{vcan}) therefore corresponds to attractive
521: interactions between particles.
522: The model exhibits a trend toward segregation,
523: which leads to a thermodynamical condensation transition
524: if $b$ is strong enough.
525: It turns out that many characteristics of this condensation transition
526: are universal~\cite{loan,evans1,wis1,cg,gross,evans2,glcond}:
527: they only depend on the asymptotic behavior of the rate
528: $u_k$ (or $v_k$) at large $k$, i.e., essentially on the value of $b$.
529: 
530: For the choice of rate~(\ref{ucan}) or~(\ref{vcan}),
531: the factor $p_k$ of the stationary-state product measure
532: (see~(\ref{pzrp}) or~(\ref{pvdef})) reads
533: \beq
534: p_k=\frac{\Gamma(b+1)\,k!}{\Gamma(k+b+1)}=b\int_0^1(1-u)^{b-1}u^k\,\d u,
535: \label{pcan}
536: \eeq
537: so that
538: \beq
539: P(z)=b\int_0^1\frac{(1-u)^{b-1}}{1-zu}\,\d u.
540: \eeq
541: The factor $p_k$ falls off as a power law with exponent $b$ at large $k$:
542: \beq
543: p_k\approx\frac{\Gamma(b+1)}{k^b}.
544: \eeq
545: The canonical ZRP has a condensation transition
546: in the thermodynamic limit whenever the first moment of the factor $p_k$,
547: \beq
548: \rho_c=\sum_{k\ge1}k\,p_k=\frac{P'(1)}{P(1)},
549: \eeq
550: is convergent.
551: We have $P(1)=b/(b-1)$ and $P'(1)=b/((b-1)(b-2))$,
552: so that the critical density $\rho_c$ is finite for $b>2$, and reads
553: \beq
554: \rho_c=\frac{P'(1)}{P(1)}=\frac{1}{b-2}.
555: \eeq
556: This critical density separates a fluid phase and a condensed phase:
557: 
558: \begin{itemize}
559: \item
560: At the {\it critical density} ($\rho=\rho_c$, $z=1$),
561: the occupation distribution reads
562: \beq
563: f_k=\frac{p_k}{P(1)}=\frac{b-1}{b}\,p_k
564: \label{fkc}
565: \eeq
566: (see~(\ref{fdef})).
567: In particular the fraction of empty sites is
568: \beq
569: f_0=\frac{b-1}{b},
570: \label{f0c}
571: \eeq
572: whereas the distribution falls off as a power law for large occupations:
573: \beq
574: f_k\approx\frac{(b-1)\Gamma(b)}{k^b}.
575: \label{fktail}
576: \eeq
577: The statics and the dynamics of the model
578: exhibit many features of critical phenomena, including scaling and universality.
579: 
580: \item
581: In the {\it fluid phase} ($\rho<\rho_c$, $z<1$),
582: the occupation distribution $f_k$ falls off exponentially.
583: 
584: \item
585: In the {\it condensed phase} $(\rho>\rho_c)$, for a large finite system,
586: the particles are arranged so as to form a uniform critical background
587: and a macroscopic condensate, typically occupying one single site
588: and consisting of $N-M\rho_c=M(\rho-\rho_c)$ excess particles.
589: \end{itemize}
590: 
591: The stationary-state measure of the asymmetric canonical target process is not
592: a product measure in the presence of interactions,
593: i.e., for any non-zero value of the parameter~$b$,
594: because the rate~(\ref{vcan}) is not of the form~(\ref{two}).
595: This absence of a product measure also holds for
596: the asymmetric target process on higher-dimensional lattices.
597: The derivation of the conditions~(\ref{m1}),~(\ref{m2})
598: given in~\cite{lux} could indeed easily be extended to the case
599: of a biased one-particle dynamics on any higher-dimensional lattice.
600: The rest of this paper is devoted to a detailed investigation of this model.
601: The study of the one-dimensional totally asymmetric model
602: is presented in Sections~\ref{1D1} and~\ref{1D2},
603: whereas Section~\ref{2D} is devoted to a maximally asymmetric form of the model
604: on the two-dimensional square lattice.
605: 
606: \section{One-dimensional target process: theoretical analysis}
607: \label{1D1}
608: 
609: In this section, we consider the target process defined by the rate~(\ref{vcan})
610: in the totally asymmetric one-dimensional case ($p=1$).
611: The situation of most physical interest will turn out to be the regime
612: of a high density.
613: 
614: We begin the analysis of the model by exploring the consequences of
615: the existence of a conserved current.
616: Consider a large finite system, with periodic boundary conditions.
617: In the stationary state,
618: the mean current of particles through the system is conserved:
619: it assumes the same value $J$ through every bond.
620: The current through the bond between sites $m$ and $m+1$
621: is given by the mean value of the corresponding rate:
622: \beq
623: J=\mean{W_{N_m,N_{m+1}}}.
624: \label{jdef}
625: \eeq
626: The existence of this conserved current is expected to ensure
627: some degree of homogeneity of the stationary state.
628: 
629: To illustrate the method, let us first consider the totally asymmetric ZRP,
630: with rate~(\ref{ucan}).
631: The current therefore reads
632: \beq
633: J=\mean{u_{N_m}}.
634: \label{jzrpdef}
635: \eeq
636: Consider a typical configuration in the condensed phase ($\rho>\rho_c$).
637: The site where the condensate is located contains a macroscopic number of
638: particles, so that~(\ref{jzrpdef}) yields $J=1$
639: for the bond to the right of the condensate,
640: up to a negligible finite-size correction of order~$1/M$.
641: For all the other bonds, the relations~(\ref{fdef}) and~(\ref{urel}) yield
642: \beq
643: J=\sum_{k\ge1}\underbrace{u_k\,f_k}_{\ds zf_{k-1}}=z,
644: \eeq
645: where $z$ is the fugacity.
646: Equating the two above expressions for the current,
647: we recover well-known results for the totally asymmetric ZRP,
648: i.e., $z=z_c=1$ and $J=1$ throughout the condensed phase.
649: 
650: Let us now turn to the totally asymmetric target process,
651: with rate~(\ref{vcan}).
652: Equation~(\ref{jdef}) for the current now reads
653: \beq
654: J=\mean{(1-\delta_{N_m,0})v_{N_{m+1}}}.
655: \label{jtpdef}
656: \eeq
657: At variance with~(\ref{jzrpdef}),
658: this expression involves the occupations of two consecutive sites.
659: 
660: The high-density regime turns out to be the situation of most physical interest.
661: In this regime, at least some of the sites must have large occupations.
662: We are therefore led to distinguish between two types of sites:
663: \begin{itemize}
664: \item
665: B-sites (B for {\it big}), whose occupation is large, of the order of $\rho$.
666: \item
667: S-sites (S for {\it small}), whose occupation is small and fluctuating.
668: \end{itemize}
669: 
670: This distinction will be kept at a heuristic level throughout the following.
671: A typical high-density configuration therefore consists of
672: four types of bonds: BB, BS, SB, and~SS.
673: The typical values of the current in each type of bond
674: obey the following inequalities in the high-density limit:
675: \beq
676: J_\SS<\matrix{J_\SB\cr J_\BS}<J_\BB.
677: \label{jineq}
678: \eeq
679: Each factor in~(\ref{jtpdef}) is indeed less than unity,
680: and approaches unity in the limit where the involved occupation goes to
681: infinity.
682: In the high-density limit we have therefore $J_\BB=1$,
683: whereas $J_\BS=\mean{v_{N_m}}$ and $J_\SB=\mean{1-\delta_{N_m,0}}$,
684: where $m$ is a typical S-site, and~$J_\SS$ is smaller
685: than the last two expressions.
686: 
687: Thanks to the inequalities~(\ref{jineq})
688: the presence of a stable isolated condensate is excluded.
689: This would indeed correspond to a current profile where the two bonds
690: on either side of the condensate carry currents $J=J_\SB$ and $J=J_\BS$
691: which are significantly higher than the background current
692: $J=J_\SS$ of all the other bonds.
693: This non-uniform current distribution with a point defect would have the effect
694: that the condensate would soon dissolve into the background.
695: The inequalities~(\ref{jineq}) actually only leave out two scenarios
696: for typical stationary-state configurations in the high-density limit.
697: Both scenarios, described below and illustrated in Figure~\ref{figunialt},
698: correspond to spatially homogeneous phases.
699: We are therefore led to predict that the asymmetric target process
700: has no condensation transition at any finite density in one dimension.
701: 
702: \begin{figure}[!tb]
703: \begin{center}
704: \includegraphics[angle=90,height=4truecm]{figuni.eps}
705: {\hskip 60pt}
706: \includegraphics[angle=90,height=4truecm]{figalt.eps}
707: \caption{\small
708: Typical configurations in the two possible scenarios
709: for the stationary state of the asymmetric one-dimensional target process
710: at high density.
711: Left: uniform configuration.
712: Right: alternating configuration.}
713: \label{figunialt}
714: \end{center}
715: \end{figure}
716: 
717: \begin{itemize}
718: \item {\it Uniform scenario.}
719: In the uniform scenario, shown in the left panel of Figure~\ref{figunialt},
720: typical configurations are entirely made of B-sites,
721: whose typical occupation is around the density $\rho$.
722: These configurations carry a current $J=J_\BB$, i.e.,
723: \beq
724: J=1
725: \eeq
726: in the high-density limit, up to a correction of order $1/\rho$.
727: 
728: \item {\it Alternating scenario.}
729: In the alternating scenario, shown in the right panel of Figure~\ref{figunialt},
730: typical configurations are alternating, i.e., they have the form BSBSBSBS...
731: 
732: The asymptotic value of the current $J=J_\BS=J_\SB$
733: through a perfect alternating structure
734: in the high-density limit can be evaluated as follows.
735: Most particles belong to B-sites, whose typical occupation is very high,
736: around $2\rho$.
737: Every B-site therefore acts as a reservoir,
738: so that the occupations of the S-sites evolve independently from each other.
739: The master equations for the occupation distribution $f_k(t)$
740: of any given S-site read
741: \beqa
742: \frad{\d f_k}{\d t}&=&f_{k+1}+v_{k-1}f_{k-1}-(1+v_k)f_k\qquad(k\ge1),
743: \nonumber\\
744: \frad{\d f_0}{\d t}&=&f_1-v_0f_0.
745: \eeqa
746: The stationary-state solution of these equations is such that
747: $f_{k+1}=v_kf_k$.
748: It is therefore proportional to the factor $p_k$.
749: The properly normalized solution is given by~(\ref{fkc}).
750: To sum up, the occupations of the S-sites are independent variables,
751: whose distribution coincides with the critical occupation distribution
752: of the dual ZRP or, equivalently, of the symmetric target process.
753: We have
754: \beq
755: J=\sum_{k\ge0}\underbrace{v_k\,f_k}_{\ds f_{k+1}}=1-f_0.
756: \eeq
757: Using (\ref{f0c}), the current in the high-density limit reads
758: \beq
759: J=\frac{1}{b}.
760: \label{jcrit}
761: \eeq
762: This alternating scenario holds a priori whenever $b>1$.
763: The latter condition corresponds to $P(1)$ being finite,
764: so that~(\ref{fkc}) is a properly normalized probability distribution.
765: \end{itemize}
766: 
767: In order to have a hint on which of the two above scenarios is preferred,
768: it is interesting to first consider the simple example of a system of two sites.
769: In this case, there is only one kind of move for the particles,
770: namely from one site to the other, and only one degree of freedom,
771: the occupation $N_1=k$ of site number~1.
772: Indeed the occupation of site number~2 reads $N_2=N-k$.
773: The stationary-state occupation distribution~$f_k$ is
774: clearly equal to the equilibrium product measure
775: \beq
776: f_k=\frac{p_kp_{N-k}}{Z_{2,N}}\approx\frac{C_N}{k^b(N-k)^b}.
777: \eeq
778: When the number $N$ of particles is large,
779: the above distribution exhibits a segregation phenomenon
780: for any positive value of $b$ (see~\cite{glcond} for a more detailed analysis,
781: including an asymptotic analysis of the amplitude $C_N$).
782: The most probable configurations are those where almost all the
783: particles are at one site, i.e., either $k\ll N$ or $N-k\ll N$.
784: This simple example confirms that the target model, just as the ZRP,
785: manifests a trend toward segregation at high density.
786: It therefore suggests that the preferred scenario
787: is that of an alternating structure.
788: 
789: This picture can be corroborated and made more quantitative
790: by means of the following dynamical stability analysis of the uniform situation.
791: For the sake of generality, in this part of the analysis
792: we deal with the partially asymmetric target process with bias $p$.
793: Consider a configuration of the uniform scenario.
794: All the sites have very high local densities $\mean{N_m}=\rho_m(t)$.
795: These local densities obey the exact rate equation
796: \beqa
797: \frac{\d\rho_m}{\d t}
798: &=&p\mean{(1-\delta_{N_{m-1},0})v_{N_m}}
799: +q\mean{(1-\delta_{N_{m+1},0})v_{N_m}}\nonumber\\
800: &-&p\mean{(1-\delta_{N_m,0})v_{N_{m+1}}}
801: -q\mean{(1-\delta_{N_m,0})v_{N_{m-1}}}.
802: \label{rhorat}
803: \eeqa
804: In the high-density regime it is legitimate
805: to simplify the above equation in several respects.
806: The probability that a site is empty is negligible,
807: whereas $\mean{v_{N_m}}\approx v_{\mean{N_m}}\approx1-b/\mean{N_m}$.
808: We are thus left with
809: \beq
810: \frac{\d\rho_m}{\d t}\approx b
811: \left(\frac{p}{\rho_{m+1}}+\frac{q}{\rho_{m-1}}-\frac{1}{\rho_m}\right).
812: \eeq
813: Let us furthermore assume that the density profile is close
814: to being constant, i.e.,
815: \beq
816: \rho_m=\rho+\dr_m,
817: \eeq
818: with $\dr_m\ll\rho$.
819: The rate equation~(\ref{rhorat}) can then be linearized as
820: \beq
821: \frac{\d\,\dr_m}{\d t}\approx\frac{b}{\rho^2}
822: \left(\dr_m-p\,\dr_{m+1}-q\,\dr_{m-1}\right).
823: \eeq
824: The component of the density profile at wavevector $K$
825: therefore grows exponentially in time as $\hdr(K)\sim\exp(\sigma(K)\,t)$,
826: where the characteristic rate $\sigma(K)$ is given by the dispersion relation
827: \beq
828: \sigma(K)=\frac{b}{\rho^2}\left(1-p\,\e^{-\i K}-q\,\e^{\i K}\right).
829: \eeq
830: The growth rate $s(K)$ is given by the real part of $\sigma(K)$, which reads
831: \beq
832: s(K)=\frac{2b}{\rho^2}\;\sin^2\frac{K}{2},
833: \label{grow}
834: \eeq
835: irrespective of the bias $p$.
836: The expression~(\ref{grow}) for the growth rate of fluctuations
837: around the uniform situation is manifestly positive
838: for all values of the wavevector $K$.
839: 
840: The uniform scenario is therefore fully linearly unstable.
841: As a consequence, the alternating scenario is the preferred one.
842: Furthermore, the alternating structure is already appearing
843: as the most favored one within the stability analysis.
844: The most unstable mode indeed corresponds to $K=\pi$,
845: i.e., an alternating density modulation of the form $\dr_m\sim(-1)^m$.
846: A uniform high-density initial configuration
847: is therefore expected to smoothly relax to an alternating one
848: by the dynamics of the asymmetric target process.
849: This observation deserves, however, to be complemented with
850: the following caveat.
851: Consider the symmetric target process, corresponding to $p=1/2$.
852: The stability analysis still leads to the expression~(\ref{grow}),
853: so that the alternating mode is still the most favored one.
854: On the other hand, the stationary state is known to be described
855: by a product measure, and especially to have a single condensate.
856: This suggests that the relaxation dynamics of a uniform
857: high-density initial configuration will exhibit two stages:
858: first, a rather fast relaxation to an intermediate structure with
859: alternating fluctuations,
860: then, a coarsening evolution of the system toward its true fate,
861: by the merging of the excess particles into fewer and fewer condensate
862: precursors.
863: This two-stage relaxation should also hold for the dual ZRP
864: with a uniform high-density initial configuration.
865: The dynamical stability analysis of the uniform situation
866: in the ZRP indeed yields exactly the same expression~(\ref{grow}),
867: again irrespective of the bias $p$.
868: 
869: To close up, we mention that the expression~(\ref{grow})
870: also yields some hints on the time scales involved in the model.
871: The characteristic time of the most unstable mode, $T_\loc=1/s(\pi)$,
872: gives an estimate of the local relaxation time of the alternating structure,
873: at the spatial scale of two consecutive sites.
874: On the other hand,
875: the long-distance behavior of the dynamics is {\it antidiffusive}.
876: We have indeed formally $s(K)\approx -DK^2$ as $K\to0$,
877: with a small negative diffusion coefficient $D=-b/(2\rho^2)$.
878: For a large but finite system made of $M$ sites,
879: with periodic boundary conditions,
880: the longest characteristic time, $T_\gl=1/s(2\pi/M)$,
881: gives an estimate of the global relaxation time of the structure as a whole.
882: In the high-density regime, the characteristic times thus defined scale as
883: \beq
884: T_\loc\approx\frac{\rho^2}{2b},\qquad T_\gl\approx\frac{\rho^2M^2}{2\pi^2b}.
885: \label{times}
886: \eeq
887: The predicted divergence of both characteristic times with density
888: provides an a posteriori confirmation that the high-density regime
889: is indeed the most interesting one.
890: 
891: \section{One-dimensional target process: numerical results}
892: \label{1D2}
893: 
894: In this section we complement our analysis
895: of the target process in the asymmetric one-dimensional case,
896: by means of numerical simulations and scaling arguments.
897: For definiteness we restrict the study to the totally asymmetric
898: situation ($p=1$).
899: Furthermore, we set once for all $b=4$.
900: We successively consider features of the transient dynamics,
901: of the stationary-state measure, and of the stationary-state dynamics.
902: The main focus is on the scaling behavior of quantities of interest
903: in the high-density regime.
904: Two types of initial conditions are considered:
905: \begin{itemize}
906: \item Deterministic initial condition:
907: the occupations of all the sites are set equal to $N_m=\rho$
908: (provided the density $\rho$ is an integer).
909: \item Random initial condition:
910: the occupations $N_m$ are drawn independently at random
911: from the geometric distribution~(\ref{fexpo}) at density $\rho$.
912: \end{itemize}
913: 
914: We start by investigating the early stage of the dynamics,
915: where relatively fast rearrangements of particles
916: bring the system to a locally stationary state.
917: This stage of the relaxation dynamics can be monitored
918: by means of a local probe at one site.
919: We choose the reduced second moment of the occupations:
920: \beq
921: K(t)=\frac{1}{\rho^2}\sum_{k\ge0}k^2f_k(t),
922: \label{kdef}
923: \eeq
924: where $f_k(t)$ is the time-dependent distribution of the occupations.
925: The initial values of this quantity are
926: $K(0)=1$ for the deterministic initial condition,
927: and $K(0)=(2\rho+1)/\rho$ for the random one.
928: 
929: Figure~\ref{figkt} shows a plot of $K(t)$,
930: measured by means of a numerical simulation, against time $t$,
931: for $\rho=10$, and a deterministic and a random initial condition.
932: Each series of data are obtained by averaging over sufficiently many
933: histories in order to obtain a smooth signal
934: ($10^4$ histories of a system of $10^4$ sites in this case).
935: The data exhibit a rather fast rise from their initial values,
936: and converge to a common limiting value, $K\approx2.82$,
937: extracted from data for much longer times, and shown as a dashed line.
938: The data for the deterministic initial condition (lower curve)
939: increase as a monotonic function of time and present a rather sharp shoulder,
940: whereas those for the random initial condition (upper curve)
941: exhibit a non-monotonic behavior with a very flat maximum.
942: 
943: \begin{figure}[!tb]
944: \begin{center}
945: \includegraphics[angle=90,height=6truecm]{figkt.eps}
946: \caption{\small
947: Plot of the reduced second moment $K(t)$ of the occupation distribution
948: in the one-dimensional fully asymmetric target process against time~$t$,
949: for $b=4$ and $\rho=10$.
950: Lower curve: deterministic initial condition.
951: Upper curve: random initial condition.
952: Horizontal dashed line: common limiting value $K\approx2.82$.}
953: \label{figkt}
954: \end{center}
955: \end{figure}
956: 
957: The characteristic time of the rise observed in $K(t)$
958: gives a measure of the local relaxation time.
959: More precisely, we define the local time~$T_\loc$
960: by the condition that~$K(t)$ is near the middle of its rise,
961: i.e., $K(T_\loc)=2$ for the deterministic initial condition,
962: and $K(T_\loc)=2.5$ for the random initial condition.
963: Figure~\ref{figtloc} shows a plot of the numerical values
964: of the local time so defined, divided by $\rho$, against $\rho$,
965: for both types of initial conditions.
966: Here and in subsequent figures,
967: statistical errors are comparable to the symbol size.
968: The least-squares fits of the two series of data suggest a growth of the form
969: \beq
970: T_\loc\approx A\rho^2+B\rho
971: \label{tlinquad}
972: \eeq
973: in the high-density regime of interest.
974: The amplitudes~$A$ and $B$ depend on the initial condition.
975: The least-squares fits shown on the plot
976: yield $A\approx0.095$ for a deterministic initial condition
977: and $A\approx0.039$ for a random initial condition.
978: These numbers, and especially the first one,
979: are comparable to the rough estimate coming from~(\ref{times}),
980: i.e., $A=1/(2b)=1/8=0.125$.
981: 
982: \begin{figure}[!tb]
983: \begin{center}
984: \includegraphics[angle=90,height=6truecm]{figtloc.eps}
985: \caption{\small
986: Plot of the local relaxation time $T_\loc$, divided by $\rho$,
987: against density~$\rho$.
988: Upper data (empty symbols): deterministic initial condition.
989: Lower data (full symbols): random initial condition.
990: Straight lines: least-squares fits with respective slopes 0.095 and 0.039.}
991: \label{figtloc}
992: \end{center}
993: \end{figure}
994: 
995: The initial condition is irrelevant
996: for what concerns later stages of the dynamics,
997: which correspond to the emergence of global features of the stationary state.
998: Hereafter we choose to work with a random initial condition.
999: Before we turn to an analysis of these late stages,
1000: it is worth taking a glance at the spatial structure
1001: of the stationary state in the high-density regime.
1002: 
1003: Figure~\ref{figtrack} shows a typical stationary
1004: occupation profile for $\rho=50$.
1005: In order to better reveal the alternating structure predicted
1006: in Section~\ref{1D1},
1007: we have plotted $(-1)^mN_m$ against the position $m$ of the site.
1008: The alternating structure clearly emerges from this representation:
1009: domains where the signal is positive (resp.~negative)
1010: correspond to domains where the B-sites are the even (resp.~odd) sites.
1011: 
1012: \begin{figure}[!tb]
1013: \begin{center}
1014: \includegraphics[angle=90,height=6truecm]{figtrack.eps}
1015: \caption{\small
1016: Plot of $(-1)^mN_m$ against the position $m$ of the site,
1017: for a typical stationary-state configuration for $\rho=50$,
1018: emphasizing the alternating structure.}
1019: \label{figtrack}
1020: \end{center}
1021: \end{figure}
1022: 
1023: In order to turn this observation into a quantitative measurement,
1024: let us introduce the concept of {\it defects}.
1025: Roughly speaking, a defect is a site around which the structure is not perfectly
1026: of the alternating form BSBSBSBS...
1027: and a domain is any stretch between two consecutive defects.
1028: More precisely, the site~$m$ is considered as a defect whenever
1029: its occupation $N_m$ is neither a maximum nor a minimum of the density profile.
1030: Equivalently,~$N_m$ is between $N_{m-1}$ and $N_{m+1}$,
1031: i.e., the product $(N_m-N_{m-1})(N_m-N_{m+1})$ is negative.
1032: This definition pinpoints 24 defects
1033: in the configuration shown in Figure~\ref{figtrack}.
1034: This number is slightly above the number of domains visible with
1035: the naked eye, i.e.,~18, because some of the domains are microscopic.
1036: Most defects can be viewed either as a BB sequence or as an SS sequence.
1037: We shall return later on to the dynamics of these defects.
1038: The density of defects $R$, i.e., the mean number of defects
1039: per unit length, reads
1040: \beq
1041: R=\bigmean{\Theta\Bigl(-(N_m-N_{m-1})(N_m-N_{m+1})\Bigr)},
1042: \eeq
1043: where the Heaviside step function on the integers is defined as
1044: \beq
1045: \Theta(n)=\left\{\matrix{
1046: 1\quad\hbox{for}\;\;n>0,\hfill\cr
1047: 0\quad\hbox{for}\;\;n\le0.}\right.%}
1048: \eeq
1049: The inverse of the density of defects,
1050: \beq
1051: \xi=\frac{1}{R},
1052: \eeq
1053: is interpreted as the mean size of a domain or, equivalently,
1054: as the {\it coherence length} of the alternating structure.
1055: 
1056: Let us now return to the late stages of the dynamics,
1057: starting from a random initial condition.
1058: In order to characterize the growth of the alternating structure,
1059: we have measured the time dependence of the mean domain size $\xi(t)=1/R(t)$.
1060: Figure~\ref{figd} shows a plot of~$\xi(t)$,
1061: for various values of the density $\rho$.
1062: For a random initial condition, we have $\xi(0)=3$.
1063: Consider indeed the initial values of $N_{m-1}$, $N_m$, and $N_{m+1}$.
1064: The probability that these three independent random numbers
1065: obey either of the inequalities $N_{m-1}<N_m<N_{m+1}$
1066: or $N_{m-1}>N_m>N_{m+1}$ is equal to 1/6 in the high-density regime
1067: (neglecting the fact that these are integer variables,
1068: which may coincide with a small but nonzero probability).
1069: Hence $R(0)=2\times1/6=1/3$ and $\xi(0)=1/R(0)=3$.
1070: The data are plotted against the reduced time variable $(t/T_\loc)^{1/2}$,
1071: for each value of the density~$\rho$,
1072: where $T_\loc$ is taken from the lower data of Figure~\ref{figtloc}.
1073: The observed common initial linear behavior,
1074: shown as a dashed straight line starting from the known value $\xi(0)=3$,
1075: demonstrates that the mean domain size grows according to the coarsening~law
1076: \beq
1077: \xi(t)\sim(t/T_\loc)^{1/2},
1078: \label{coa}
1079: \eeq
1080: before it saturates to a density-dependent
1081: stationary-state value, simply denoted by $\xi$.
1082: The duration of the coarsening process, before the stationary state is reached,
1083: defines the global relaxation time $T_\gl$ of the problem.
1084: By inverting the relation~(\ref{coa}),
1085: we predict that the latter time grows as $T_\gl\sim T_\loc\,\xi^2$.
1086: This relation between $T_\loc$ and $T_\gl$ is in agreement with~(\ref{times}),
1087: where the stationary-state mean domain size $\xi$ plays the role
1088: of the system size $M$.
1089: 
1090: \begin{figure}[!tb]
1091: \begin{center}
1092: \includegraphics[angle=90,height=6truecm]{figd.eps}
1093: \caption{\small
1094: Plot of the mean domain size $\xi(t)$
1095: against the reduced time $(t/T_\loc)^{1/2}$,
1096: for various values of the density $\rho$, indicated on the curves.
1097: The dashed straight line starting from $\xi(0)=3$
1098: illustrates the coarsening law~(\ref{coa}).}
1099: \label{figd}
1100: \end{center}
1101: \end{figure}
1102: 
1103: We now investigate a few characteristic features
1104: of the nonequilibrium stationary state of the asymmetric target process,
1105: emphasizing their scaling behavior at high density.
1106: We start with the mean domain size~$\xi$.
1107: Its scaling behavior at high density can be predicted by the following argument.
1108: Roughly speaking, a defect can be thought of as a B-site whose occupation
1109: is accidentally as small as that of an S-site, i.e., finite.
1110: Anticipating the scaling law~(\ref{bsca}),
1111: and assuming a linear rise for the scaling function~$F(x)$,
1112: we find that the probability of such an event scales as $1/\rho^2$.
1113: This argument leads to an asymptotic
1114: quadratic growth of the mean domain size of the~form
1115: \beq
1116: \xi\approx\Xi\,\rho^2.
1117: \label{xiquad}
1118: \eeq
1119: 
1120: \begin{figure}[!tb]
1121: \begin{center}
1122: \includegraphics[angle=90,height=6truecm]{figxi.eps}
1123: \caption{\small
1124: Plot of the mean domain size $\xi$ in the stationary state
1125: against density~$\rho$.
1126: Full line: second-degree polynomial fit to the data
1127: with leading coefficient $\Xi=416\times10^{-6}$.}
1128: \label{figxi}
1129: \end{center}
1130: \end{figure}
1131: 
1132: Figure~\ref{figxi} shows a plot of the stationary-state value
1133: of $\xi$ against density $\rho$, for densities up to $\rho=200$.
1134: The second-degree polynomial fit to the data is compatible with
1135: our expectation~(\ref{xiquad}).
1136: The numerical value for the prefactor, $\Xi\approx4\times10^{-4}$,
1137: is however found to be very small.
1138: Partly as a consequence of this smallness,
1139: the data exhibit large corrections to the above asymptotic law
1140: for values of the density accessible to numerical simulations.
1141: At variance with the case of $T_\loc$, plotted in Figure~\ref{figtloc},
1142: we found no way to unambiguously characterize these corrections.
1143: The quadratic growth~(\ref{xiquad}) of~$\xi$
1144: corresponds to a very fast growth of the global relaxation time:
1145: \beq
1146: T_\gl\sim T_\loc\,\xi^2\sim\rho^6.
1147: \eeq
1148: 
1149: \begin{figure}[!tb]
1150: \begin{center}
1151: \includegraphics[angle=90,height=6truecm]{figj.eps}
1152: {\hskip 15pt}
1153: \includegraphics[angle=90,height=6truecm]{figk.eps}
1154: \caption{\small
1155: Plot of stationary values of the current $J$ (left) and of the
1156: reduced second moment~$K$ (right),
1157: against the stationary density of defects $R=1/\xi$.
1158: Symbols: data for densities ranging from $\rho=30$ to 200.
1159: Full lines: polynomial fits, of degree 2 for $J$ and 3 for $K$,
1160: yielding the extrapolated values $J_\infty\approx0.250$
1161: and $K_\infty\approx2.98$.
1162: Dashed line: constrained polynomial fit of degree 3 for $K$,
1163: imposing $K_\infty=3$ (see text).}
1164: \label{figjk}
1165: \end{center}
1166: \end{figure}
1167: 
1168: Figure~\ref{figjk} shows a plot of the stationary-state values
1169: of the current $J$ and of the reduced second moment $K$
1170: of the occupation distribution,
1171: against the stationary-state density of defects $R=1/\xi$.
1172: The second-degree polynomial fit for $J$ yields $J_\infty\approx0.250$,
1173: in excellent quantitative agreement with
1174: the limiting value $J=1/b=1/4$ predicted in~(\ref{jcrit}).
1175: A third-degree polynomial fit for $K$ yields the
1176: limiting value $K_\infty\approx2.98$.
1177: This number is very close to the value $K_\infty=3$ corresponding to the
1178: trial scaling function $F_0(x)$ given in~(\ref{ftrial}).
1179: This proximity suggests that the value $K_\infty=3$ could be exact.
1180: Imposing the constraint $K_\infty=3$ indeed hardly alters
1181: the quality of the fit (dashed line).
1182: 
1183: \begin{figure}[!tb]
1184: \begin{center}
1185: \includegraphics[angle=90,height=6truecm]{figf1.eps}
1186: {\hskip 15pt}
1187: \includegraphics[angle=90,height=6truecm]{figf2.eps}
1188: \caption{\small
1189: Stationary-state occupation distribution $f_k$.
1190: The left panel shows data for $k\ll\rho$ (S-sites).
1191: Empty symbols: plot of $\ln f_k$ against $\ln(k+1)$
1192: for $\rho=50$ (upper data) and 100 (lower data).
1193: Full symbols (labeled ZRP): plot of $\ln (f_k/2)$,
1194: where $f_k$ is the critical distribution of the dual ZRP.
1195: The right panel shows data for $k$ comparable to the density $\rho$ (B-sites).
1196: Full lines: scaling plots of the product $\rho f_k$ against $k/\rho$
1197: for $\rho=50$ (upper data) and 100 (lower data).
1198: Symbols: minimum value of $f_k$.
1199: Dashed line: trial scaling function $F_0(x)$ defined in~(\ref{ftrial}).}
1200: \label{figf}
1201: \end{center}
1202: \end{figure}
1203: 
1204: In complete agreement with the alternating scenario
1205: depicted in Section~\ref{1D1},
1206: the occupation distribution $f_k$ in the stationary state
1207: consists of two distinct components with equal weights,
1208: which respectively describe S-sites and B-sites.
1209: These components are emphasized in Figure~\ref{figf},
1210: where the data for $f_k$ for the same values of the density, $\rho=50$ and 100,
1211: are plotted in two different ways.
1212: The left panel shows a plot of $\ln f_k$ against $\ln(k+1)$
1213: for moderate values of the occupation $k$, up to~20.
1214: These are, roughly speaking, the S-sites.
1215: The data for the smaller values of~$k$
1216: are very close to $1/2$ times the critical occupation distribution
1217: of the dual ZRP or, equivalently, of the symmetric target process,
1218: given by~(\ref{fkc}), also shown on the plot (full symbols).
1219: The range of values of $k$ over which the agreement holds
1220: is observed to get larger for larger densities.
1221: This is a convincing confirmation of the prediction made
1222: in Section~\ref{1D1} that the S-sites are critical in the stationary state.
1223: The right panel shows a scaling plot of the product $\rho f_k$ against $k/\rho$,
1224: for larger values of $k$, comparable to the density $\rho$.
1225: These are, roughly speaking, the B-sites.
1226: The data exhibit a scaling law of the form
1227: \beq
1228: f_k\approx\frac{1}{\rho}\;F\!\left(\frac{k}{\rho}\right).
1229: \label{bsca}
1230: \eeq
1231: The scaling function $F(x)$ obeys the sum rules
1232: \beq
1233: \int_0^\infty F(x)\,\d x=1/2,\quad\int_0^\infty xF(x)\,\d x=1,\quad
1234: \int_0^\infty x^2F(x)\,\d x=K_\infty.
1235: \eeq
1236: The first two equalities express that the fraction of B-sites is $1/2$
1237: and that their mean occupation is $2\rho$, whereas
1238: the third one is a rewriting of the definition of~$K_\infty$.
1239: The scaling function is observed to be rather uniformly well approximated by
1240: the trial scaling function
1241: \beq
1242: F_0(x)=\frac12\,x\,\e^{-x},
1243: \label{ftrial}
1244: \eeq
1245: shown on the right panel of Figure~\ref{figf} as a dashed line.
1246: The scaling function~(\ref{ftrial}) corresponds to $K_\infty=3$.
1247: The closeness of this number to the extrapolated value $K_\infty\approx2.98$
1248: opens up the possibility that the scaling function $F(x)$ is exactly given
1249: by $F_0(x)$.
1250: In any case, the linear rise of the trial function $F_0(x)$
1251: seems to be shared by the true scaling function $F(x)$.
1252: 
1253: Both components of the occupation distribution shown in Figure~\ref{figf},
1254: respectively corresponding to S-sites and B-sites,
1255: are separated by a minimum in the occupation distribution.
1256: In the high-density regime
1257: this minimum takes place for a crossover occupation $k_\star$
1258: such that the estimates $1/k^b$ (see~(\ref{fktail})) and $k/\rho^2$
1259: (assuming a linear rise for the scaling function $F(x)$) are comparable.
1260: We thus obtain
1261: \beq
1262: k_\star\sim\rho^{2/(b+1)},\qquad f_{k_\star}\sim\rho^{-2b/(b+1)}.
1263: \label{xover}
1264: \eeq
1265: These estimates make sense as soon as the crossover occupation obeys
1266: $k_\star\ll\rho$.
1267: We thus recover the condition~$b>1$ for the validity
1268: of the alternating scenario.
1269: The mean occupation $\rho_S$ of the S-sites has the finite asymptotic value
1270: $\rho_c$ for $b>2$,
1271: whereas it scales as $\rho_S\sim k_\star^{2-b}\sim\rho^{2(2-b)/(b+1)}$
1272: in the high-density regime for $1<b<2$.
1273: 
1274: We close up this section
1275: with an investigation of the stationary-state dynamics of defects.
1276: For a large but finite density $\rho$,
1277: there is a small density of defects $R=1/\xi\sim1/\rho^2$ (see~(\ref{xiquad})).
1278: These defects cannot stay immobile.
1279: 
1280: \begin{itemize}
1281: \item
1282: Consider indeed a BB defect, made of two consecutive B-sites.
1283: The current between the two B-sites, $J_\BB=1$, exceeds the mean current
1284: $J=1/b$ through the system.
1285: As a consequence, particles flow from the left B-site of the defect
1286: into the right one at a rate $\omega_\BB=J_\BB-J=(b-1)/b$.
1287: After a time of the order of
1288: \beq
1289: \tau_\BB\approx\frac{2\rho}{\omega_\BB}=\frac{2b\rho}{b-1},
1290: \eeq
1291: the left B-site is emptied.
1292: This is the first reaction of~(\ref{bbss}).
1293: An SS defect is thus formed one site to the left of the original BB defect.
1294: 
1295: \item
1296: Consider now an SS defect, made of two consecutive S-sites.
1297: The current between the two S-sites, $J_\SS\approx1/b^2$,
1298: is smaller than the mean current $J$.
1299: As a consequence, particles flow from the B-site to the left of the defect
1300: into the left S-site at a rate $\omega_\SS=J-J_\SS\approx(b-1)/b^2$.
1301: The left S-site is thus soon (i.e., after a time which does not
1302: grow proportionally to $\rho$) turned to a B-site.
1303: This is the second reaction of~(\ref{bbss}).
1304: A BB defect is thus formed one site to the left of the original~SS defect.
1305: \end{itemize}
1306: 
1307: The discussion can be summarized in the form of the following reactions
1308: \beq
1309: \BB\to\SB,\qquad\SS\to\BS.
1310: \label{bbss}
1311: \eeq
1312: The typical history of a single defect therefore looks as follows,
1313: where time runs from bottom to top,
1314: for the sake of consistency with Figure~\ref{figplot}:
1315: \[
1316: t\uparrow\quad\matrix{
1317: \BS\BS\bfBB\SB\SB\SB\SB\SB\SB\SB\cr
1318: \BS\BS{\rm B}\bfSS{\rm B}\SB\SB\SB\SB\SB\SB\cr
1319: \BS\BS\BS\bfBB\SB\SB\SB\SB\SB\SB\cr
1320: \BS\BS\BS{\rm B}\bfSS{\rm B}\SB\SB\SB\SB\SB\cr
1321: \BS\BS\BS\BS\bfBB\SB\SB\SB\SB\SB\cr
1322: \BS\BS\BS\BS{\rm B}\bfSS{\rm B}\SB\SB\SB\SB\cr
1323: \BS\BS\BS\BS\BS\bfBB\SB\SB\SB\SB\cr
1324: \BS\BS\BS\BS\BS{\rm B}\bfSS{\rm B}\SB\SB\SB\cr
1325: n\to}
1326: \]
1327: 
1328: \begin{figure}[!tb]
1329: \begin{center}
1330: \includegraphics[angle=90,height=6truecm]{figplot.eps}
1331: \caption{\small
1332: Space-time plot of the stationary domain pattern for $\rho=100$.
1333: Black (resp.~white) areas show the regions of space-time
1334: where the S-sites are the odd (resp.~the even) sites.
1335: The slope of the straight line in the right part of the plot
1336: yields $V\approx-0.0072$.}
1337: \label{figplot}
1338: \end{center}
1339: \end{figure}
1340: 
1341: The time it takes for a defect to move two sites to the left
1342: is therefore equal to $\tau_\BB$ on average.
1343: As a consequence, the pattern of defects and domains is
1344: advected with an upstream (negative) velocity $V\approx-2/\tau_\BB$, i.e.,
1345: \beq
1346: V\approx-\frac{b-1}{b\rho}.
1347: \label{vbal}
1348: \eeq
1349: This scenario is confirmed by Figure~\ref{figplot},
1350: showing a space-time plot of the stationary dynamics
1351: of the domain pattern for $\rho=100$.
1352: Equation~(\ref{vbal}) predicts $V\approx-0.0075$,
1353: in reasonably good agreement with the observed value $V\approx-0.0072$.
1354: The plot shows that the whole advected domain pattern
1355: behaves more or less as a rigid body over spatial scales
1356: much larger than the mean domain size $\xi$.
1357: 
1358: \section{Two-dimensional target process}
1359: \label{2D}
1360: 
1361: We now consider the canonical target process
1362: defined by the rate~(\ref{vcan}), on the square lattice
1363: with unit vectors $\bfe_1$, $\bfe_2$.
1364: In order to have a genuine two-dimensional model
1365: and to maximize the asymmetry, we choose the following rule:
1366: particles hop either East (displacement $\bfe_1$)
1367: or North (displacement $\bfe_2$) with equal probabilities.
1368: In other words, if the departure site is $\bfd=(m,n)$,
1369: the arrival site is chosen to be either $\bfa=(m+1,n)=\bfd+\bfe_1$,
1370: or $\bfa=(m,n+1)=\bfd+\bfe_2$, with probability $1/2$.
1371: The bias, i.e., the mean displacement proposed
1372: to a particle, $\bfb=(\bfe_1+\bfe_2)/2$, is along the North-East direction.
1373: 
1374: \subsection{Heuristic argument}
1375: 
1376: In the one-dimensional situation, the existence of a conserved current
1377: was instrumental in order to discriminate between possible scenarios
1378: for the stationary state of the model.
1379: In the present situation, however, the current $\J$ is a two-dimensional
1380: vector.
1381: The condition that~$\J$ be conserved in the stationary state
1382: is less stringent than in the one-dimensional case.
1383: On spatial scales much larger than the lattice spacing,
1384: it is reasonable to use the continuum formalism.
1385: Within this framework, the conservation law reads $\nabla\cdot\J=0$.
1386: The symmetry of the dynamical rules implies that $\J$
1387: is aligned with the bias $\bfb$.
1388: Current lines are therefore parallel straight lines along this direction.
1389: The conservation law implies that the magnitude of the current
1390: is constant along each current line,
1391: but may well vary in the transversal direction from one current line to another.
1392: In particular the existence of an {\it extended condensate} is allowed,
1393: in the form of a one-dimensional structure
1394: elongated along the direction of the bias.
1395: This opens up the possibility of having an unconventional
1396: type of condensation transition.
1397: Such a phenomenon is indeed observed in the numerical simulations described
1398: hereafter.
1399: 
1400: \subsection{Numerical results}
1401: 
1402: We choose once for all the value $b=4$ in the numerical simulations
1403: of the two-dimensional target process defined above.
1404: 
1405: In analogy with the one-dimensional case, we begin with the transient dynamics,
1406: starting from a random initial condition.
1407: We monitor the local relaxation by means of the reduced second moment $K(t)$
1408: of the occupation distribution, introduced in~(\ref{kdef}).
1409: Figure~\ref{fig2kt} shows a plot of $K(t)$ against time~$t$.
1410: For all values of the density~$\rho$, $K(t)$ is an increasing function of time,
1411: starting from its initial value $K(0)=(2\rho+1)/\rho$.
1412: The most significant feature to be observed on the data is the following:
1413: $K(t)$ saturates to a finite limiting value, denoted $K$,
1414: for values of the density up to the threshold value $\rho_0\approx6$.
1415: We have~$K_0\approx4.8$ for $\rho=\rho_0$.
1416: To the contrary, $K(t)$ grows indefinitely for larger values of the density.
1417: The right panel of Figure~\ref{fig2kt} demonstrates that
1418: the asymptotic growth law of~$K(t)$ for all~$\rho>\rho_0$ is of the form
1419: \beq
1420: K(t)\approx\frac{Ct}{\rho},
1421: \label{asylaw}
1422: \eeq
1423: where $C\approx9\times10^{-3}$ is taken from the slope
1424: of the parallel dashed lines.
1425: 
1426: \begin{figure}[!tb]
1427: \begin{center}
1428: \includegraphics[angle=90,height=6truecm]{fig2kt.eps}
1429: {\hskip 15pt}
1430: \includegraphics[angle=90,height=6truecm]{fig2klong.eps}
1431: \caption{\small
1432: Plot of the reduced second moment $K(t)$ of the occupation distribution
1433: in the two-dimensional target process against time $t$,
1434: for various values of the density $\rho$, indicated on the curves.
1435: Left: data for moderate values of time.
1436: Right: data for longer values of time, multiplied by density $\rho$.
1437: The parallel straight dashed lines, meant as a guide to the eye,
1438: have a slope $9.1\times10^{-3}$.}
1439: \label{fig2kt}
1440: \end{center}
1441: \end{figure}
1442: 
1443: Pursuing along the lines of our investigation of the one-dimensional case,
1444: we introduce a local relaxation time~$T_\loc$,
1445: defined by the condition $K(T_\loc)=4$.
1446: Figure~\ref{fig2tloc} shows a plot of $T_\loc$ so defined,
1447: divided by~$\rho$, against~$\rho$.
1448: The least-squares fit suggests a growth of the form~(\ref{tlinquad}),
1449: with $A\approx0.63$, in a wide range of values of the density.
1450: Notice that the local relaxation time exhibits no visible singularity
1451: at the threshold density $\rho_0$.
1452: 
1453: \begin{figure}[!tb]
1454: \begin{center}
1455: \includegraphics[angle=90,height=6truecm]{fig2tloc.eps}
1456: \caption{\small
1457: Plot of the local relaxation time $T_\loc$, divided by $\rho$,
1458: against density~$\rho$.
1459: Full straight line: least-squares fit with slope $A\approx0.63$.}
1460: \label{fig2tloc}
1461: \end{center}
1462: \end{figure}
1463: 
1464: We now turn to a more accurate determination of the threshold density $\rho_0$,
1465: using the pre-asymptotic growth of $K(t)$ in the intermediate time regime,
1466: where $K(t)$ has already departed in a significant way
1467: from its plateau value~$K_0$,
1468: but not yet reached the asymptotic linear growth~(\ref{asylaw}).
1469: The right panel of Figure~\ref{fig2kt} demonstrates
1470: that this regime lasts longer and longer
1471: as the threshold density $\rho_0$ is approached.
1472: This observation is turned to a quantitative measurement
1473: by defining the global relaxation time $T_\gl$
1474: by the condition $K(T_\gl)=K_0+\Delta K$,
1475: where we set $K_0=4.8$, whereas the choice $\Delta K=50/\rho$
1476: incorporates the form~(\ref{asylaw}) of the asymptotic growth law.
1477: Figure~\ref{fig2tg} shows a plot of the reciprocal of the global time $T_\gl$
1478: thus defined, against density.
1479: The data convincingly demonstrate that the global time diverges at some
1480: non-trivial threshold density~$\rho_0$.
1481: A crossover of the data toward another type of asymptotic behavior
1482: indeed seems extremely improbable,
1483: in view of the accuracy of the available data.
1484: The second-degree polynomial fit to the data shown on the plot
1485: provides a rather accurate determination of the threshold density,
1486: \beq
1487: \rho_0=6.0\pm0.1,
1488: \label{rhoc}
1489: \eeq
1490: as well as an evidence that the global time
1491: diverges linearly as the threshold density is approached from above, as
1492: \beq
1493: T_\gl\approx T_0\;\frac{\rho_0}{\rho-\rho_0},
1494: \label{tgldiv}
1495: \eeq
1496: with a rather large prefactor $T_0\approx28\,000$.
1497: 
1498: \begin{figure}[!tb]
1499: \begin{center}
1500: \includegraphics[angle=90,height=6truecm]{fig2tg.eps}
1501: \caption{\small
1502: Plot of the reciprocal of the global time $T_\gl$ against density $\rho$.
1503: Full line: second-degree polynomial fit yielding $\rho_0=6.0\pm0.1$.}
1504: \label{fig2tg}
1505: \end{center}
1506: \end{figure}
1507: 
1508: The threshold density $\rho_0$ is the maximal density for which a homogeneous
1509: fluid phase is stable.
1510: At variance with the critical density $\rho_c$ of the symmetric target process
1511: and of the dual ZRP, $\rho_0$ rather appears as a dynamical threshold.
1512: This viewpoint is corroborated by the fact
1513: that the stationary state at density~$\rho_0$ does not exhibit
1514: any critical feature.
1515: The distribution of the site occupations at the threshold density,
1516: shown in Figure~\ref{fig2f}, has an exponential fall-off
1517: of the form $f_k\sim\exp(-\mu k)$, with $\mu\approx0.06$,
1518: at least in the accessible range of values of the occupation.
1519: We checked that the data are not affected
1520: in an appreciable way by finite-size effects in the range considered.
1521: The critical occupation distribution~(\ref{fkc}) of the dual ZRP
1522: is shown on the same plot as a comparison.
1523: The latter distribution has a much smaller density $\rho_c=1/2$,
1524: some 12 times smaller than the observed threshold density~(\ref{rhoc})
1525: of the asymmetric two-dimensional model, but a slower power-law fall-off,
1526: so that the distributions eventually cross each other.
1527: 
1528: \begin{figure}[!tb]
1529: \begin{center}
1530: \includegraphics[angle=90,height=6truecm]{fig2f.eps}
1531: \caption{\small
1532: Logarithmic plot of occupation distributions.
1533: Upper curve: fully asymmetric target process in two dimensions at its threshold
1534: density $\rho_0$.
1535: The dashed line, meant as a guide to the eye, has a slope $-\mu\approx-0.06$.
1536: Lower curve: critical occupation distribution of the dual ZRP,
1537: given by~(\ref{fkc}).}
1538: \label{fig2f}
1539: \end{center}
1540: \end{figure}
1541: 
1542: The progressive emergence of highly occupied coherent structures,
1543: which are strongly elongated along the direction of the bias,
1544: is illustrated in Figure~\ref{fig2plot},
1545: showing two snapshots of the coarsening regime
1546: of a sample of size $100\times100$ at density $\rho=20$.
1547: The filled symbols show the 1\% most occupied sites.
1548: The visible structures clearly are precursors of the extended condensate
1549: mentioned above.
1550: The background density of the fluid phase besides these structures
1551: is found to be much smaller than the threshold density~$\rho_0$,
1552: and comparable to the critical density $\rho_c=1/2$ of the dual ZRP.
1553: We shall return to this point in more detail below (see Figure~\ref{figrhobg}).
1554: 
1555: \begin{figure}[!tb]
1556: \begin{center}
1557: \includegraphics[angle=90,height=6truecm]{fig2plot3.eps}
1558: {\hskip 15pt}
1559: \includegraphics[angle=90,height=6truecm]{fig2plot4.eps}
1560: \caption{\small
1561: Plots of the 100 (i.e.,~1\%) most occupied sites
1562: of a sample of size $100\times100$ at density $\rho=20$
1563: in the coarsening regime.
1564: Left: $t=5\times10^4$.
1565: Right: $t=10^5$.}
1566: \label{fig2plot}
1567: \end{center}
1568: \end{figure}
1569: 
1570: We now study the typical characteristic sizes (width and height)
1571: of the extended condensate in the stationary state of a finite system,
1572: and of its precursors in the coarsening regime of an infinite system.
1573: The width $W$ of a condensate is defined as the number of sites
1574: which take part in the condensate,
1575: whereas its height $H$ is the mean number of particles per site
1576: in the condensate, so that the product~$HW$
1577: gives a measure of the number of particles involved in the condensate.
1578: We first consider the coarsening regime of an infinite system.
1579: The condensate precursors shown in Figure~\ref{fig2plot}
1580: are expected to be characterized by a typical
1581: width $W(t)$ and height $H(t)$, with both scales growing with time.
1582: The contribution of these precursors
1583: to the reduced second moment of the occupation distribution
1584: can be checked to scale as $K(t)\sim H(t)/\rho$,
1585: irrespective of the width $W(t)$.
1586: The growth law~(\ref{asylaw}) therefore implies
1587: that the height of condensate precursors grows linearly in time, according to
1588: \beq
1589: H(t)\approx Ct,
1590: \label{ht}
1591: \eeq
1592: with $C\approx9\times10^{-3}$.
1593: This asymptotic coarsening law is expected to hold
1594: for any density~$\rho>\rho_0$.
1595: The behavior of the width $W$ of condensate precursors can only be investigated
1596: in an indirect way, by means of finite-size scaling.
1597: We therefore consider finite systems,
1598: namely square samples of linear size $L$, with periodic boundary conditions.
1599: Figure~\ref{figfss} shows numerical data concerning
1600: the stationary state of finite systems against their linear size~$L$,
1601: at fixed density $\rho=20$, well above the threshold density~$\rho_0$.
1602: The left panel shows the stationary-state value $K_L$
1603: of the reduced second moment of the occupation distribution.
1604: The right panel shows the characteristic relaxation time $T_L$,
1605: defined by the condition $K(T_L)=(K_L+K_0)/2$, again with $K_0=4.8$.
1606: The data for both quantities clearly exhibit a linear growth with the size $L$.
1607: The first of these growth laws implies $H_L\approx\rho K_L\sim L$.
1608: This is in accord with the expectation
1609: that typical stationary-state configurations
1610: have a single and roughly system-spanning extended condensate,
1611: for which $H_L\sim L$.
1612: Furthermore, as the number of particles involved in the condensate
1613: scales as $H_LW_L\sim L^2$, we have $H_L\sim W_L\sim L$.
1614: Let us now make the finite-size scaling assumption
1615: that $H(t)$ and $W(t)$ become respectively comparable to $H_L$ and $W_L$
1616: for a time $t$ comparable to the relaxation time $T_L$.
1617: This yields the scaling law $H(t)\sim t$,
1618: already known~(see~(\ref{ht})), and the prediction $W(t)\sim t$.
1619: 
1620: \begin{figure}[!tb]
1621: \begin{center}
1622: \includegraphics[angle=90,height=6truecm]{figfss1.eps}
1623: {\hskip 15pt}
1624: \includegraphics[angle=90,height=6truecm]{figfss2.eps}
1625: \caption{\small
1626: Plots of data concerning the stationary state of finite systems
1627: at fixed density $\rho=20$, against their linear size $L$.
1628: Left: stationary-state value $K_L$ of the reduced second moment
1629: of the occupation distribution.
1630: Right: relaxation time $T_L$.
1631: Full straight lines: least-squares fits with respective slopes 1.79 and
1632: 2\,660.}
1633: \label{figfss}
1634: \end{center}
1635: \end{figure}
1636: 
1637: Finally, we have also measured the background density $\rho_L^\bg$ of the fluid
1638: phase in the stationary state of finite samples at density $\rho=20$.
1639: This quantity is algorithmically defined as follows.
1640: For any intercept $k=1,\dots,L$, consider
1641: the total number of particles in the diagonal array
1642: with intercept $k$, i.e., with equation $n=m+k\ (\hbox{mod.}\ L)$:
1643: \beq
1644: \N_k=\sum_{m=1}^LN_{m,m+k}.
1645: \eeq
1646: The largest of these $L$ numbers, $\N_\max$,
1647: corresponds to the diagonal array occupied by the extended condensate.
1648: It is overwhelmingly larger than the others, as it scales as $\N_\max\sim L^2$.
1649: The other $(L-1)$ numbers $\N_k$ represent the fluid phase,
1650: and therefore scale as $\rho_L^\bg L$.
1651: We are thus naturally led to define the background density as
1652: \beq
1653: \rho_L^\bg=\frac{N-\N_\max}{L(L-1)},
1654: \eeq
1655: where $N$ is the total number of particles
1656: in the system.
1657: Figure~\ref{figrhobg} shows a plot of the stationary-state
1658: background density $\rho_L^\bg$ for finite samples of size $L$, against $1/L$.
1659: The data demonstrate that $\rho_L^\bg$ is smaller than unity,
1660: and therefore much smaller than the mean density $\rho=20$, as soon as $L\ge7$.
1661: The segregation phenomenon is therefore already fully at work
1662: for rather small system sizes.
1663: From a quantitative viewpoint, a second-degree polynomial fit to the data yields
1664: the extrapolated value $\rho^\bg\approx0.48$.
1665: The limiting value thus obtained is remarkable close
1666: to the critical density $\rho_c=1/2$ of the dual ZRP.
1667: Imposing the constraint $\rho^\bg=\rho_c=1/2$ indeed hardly changes
1668: the fit (dashed line).
1669: This agreement strongly suggests that the background fluid phase
1670: of the two-dimensional target process above its threshold density
1671: is characterized by the critical occupation distribution of the dual ZRP,
1672: just as the S-sites of the one-dimensional case in the high-density limit.
1673: As a consequence, the density $\rho^\bg$ of the fluid phase:
1674: \beq
1675: \rho^\bg=\left\{\matrix{
1676: \rho\hfill&\hbox{for}\hfill&\rho<\rho_0,\cr
1677: \rho_c\hfill&\hbox{for}\hfill&\rho>\rho_0,}\right.%}
1678: \label{rhojump}
1679: \eeq
1680: has a discontinuous jump at the threshold density $\rho_0$.
1681: 
1682: \begin{figure}[!tb]
1683: \begin{center}
1684: \includegraphics[angle=90,height=6truecm]{figrhobg.eps}
1685: \caption{\small
1686: Plot of the background density $\rho_L^\bg$
1687: in the stationary state of finite systems
1688: at fixed density $\rho=20$, against their reciprocal linear size $1/L$.
1689: Full line: second-degree polynomial fit yielding the extrapolated
1690: value $\rho^\bg\approx0.48$.
1691: Dashed line: constrained polynomial fit of degree 3 imposing
1692: $\rho^\bg=\rho_c=1/2$.}
1693: \label{figrhobg}
1694: \end{center}
1695: \end{figure}
1696: 
1697: \section{Discussion}
1698: \label{quatre}
1699: 
1700: In this work we introduced a novel example
1701: of a migration process, the target process.
1702: We then studied in detail the structure of the nonequilibrium stationary state
1703: of the asymmetric target process, the main focus being on
1704: the fate of the condensation phenomenon.
1705: 
1706: The stationary-state measures of migration processes
1707: do not have a product form in general.
1708: The {\it symmetric} target process, though,
1709: has the same stationary-state product measure as the corresponding dual ZRP.
1710: In particular, the so-called canonical target process, defined by the
1711: rate~(\ref{vcan}) dual to the ZRP with rate~(\ref{ucan}),
1712: has a continuous condensation transition
1713: at a finite critical density $\rho_c$ whenever $b>2$,
1714: with a macroscopic condensate occupying a single site
1715: for densities $\rho>\rho_c$, irrespective of the dimensionality of the system.
1716: The {\it asymmetric} target process has a fluctuating stationary state
1717: with non-trivial spatial and temporal correlations,
1718: whose qualitative features depend on the dimensionality.
1719: Our main effort in the present work consisted in characterizing
1720: this nonequilibrium stationary state,
1721: including its dependence on the dimensionality.
1722: 
1723: We have reached a complete understanding of the one-dimensional target process,
1724: especially in the high-density regime of most interest.
1725: We showed, by exploiting the existence of a conserved current,
1726: that the asymmetric canonical target process has no condensation transition,
1727: and remains homogeneous at any finite density.
1728: In the high-density regime, an alternating scenario prevails for $b>1$:
1729: typical configurations consist of long alternating sequences BSBSBSBS...
1730: of highly occupied B-sites, and less occupied S-sites, whose occupation
1731: distribution coincides with the critical distribution of the dual ZRP.
1732: The coherence length (mean domain size) of this alternating
1733: structure diverges as $\xi\sim\rho^2$.
1734: We also gave a characterization of the scaling behavior of many other
1735: quantities in the vicinity of the `infinite-density fixed point'.
1736: 
1737: For the asymmetric target process in higher dimensions,
1738: we argued that the condensate must be extended
1739: and have the form of a one-dimensional structure
1740: elongated along the direction of the bias.
1741: In the two-dimensional case, numerical simulations performed for $b=4$
1742: show that the model exhibits an unconventional condensation transition
1743: at the density $\rho_0\approx6$.
1744: This density, which is much larger than the critical density $\rho_c=1/2$
1745: of the dual ZRP, appears as a dynamical threshold:
1746: it is the maximal density at which a homogeneous fluid phase
1747: is dynamically stable.
1748: For $\rho>\rho_0$, the predicted extended condensate is observed,
1749: whereas the background fluid phase again appears as critical.
1750: This picture seems to be generic for higher-dimensional systems.
1751: Preliminary numerical simulations
1752: of the asymmetric target process on the three-dimensional cubic lattice
1753: (where the displacement is along either of the unit vectors
1754: $\bfe_1$, $\bfe_2$, $\bfe_3$ with equal probabilities),
1755: indeed show that the overall picture is quite similar
1756: to the two-dimensional one.
1757: The global relaxation time~$T_\gl$ is again found
1758: to diverge according to~(\ref{tgldiv}), with $\rho_0\approx34$ for $b=4$.
1759: 
1760: Let us mention that another mechanism
1761: leading to an extended condensate in a class of mass transport models
1762: in one dimension has been reported recently~\cite{ehm}.
1763: There, the nonequilibrium stationary-state measure is a product
1764: whose factors involve the occupations of two consecutive sites.
1765: In the condensed phase, those models exhibit an extended condensate,
1766: whose height and width scale as $H_N\sim W_N\sim N^{1/2}$
1767: for a finite system of $N$ sites.
1768: These scaling laws are formally identical to those found in the present work.
1769: 
1770: The existence of a threshold density $\rho_0$ at which
1771: the background density has a discontinuous jump (see~(\ref{rhojump}))
1772: is reminiscent of what occurs in the model studied in~\cite{zrp2},
1773: namely a ZRP with two species of particles, and with rates such that
1774: the stationary-state measure does not have a product form.
1775: When the densities $\rho^{(1)}$ and~$\rho^{(2)}$ of the two species are equal,
1776: the behavior of the system is qualitatively the same as that
1777: of the canonical ZRP (with one species).
1778: In particular the system has a continuous phase transition
1779: at some critical density $\rho_c$, from a fluid phase to a
1780: condensed phase with critical background.
1781: The general situation where the two densities are different
1782: however drastically departs from this known scenario.
1783: If either of the two densities ($\rho^{(1)}$, say) is kept fixed
1784: at a value larger than $\rho_c$, on increasing the other density $\rho^{(2)}$,
1785: the system remains homogeneous as long as $\rho^{(2)}$ is less than a
1786: threshold value $\rho^{(2)}_0$ which depends on~$\rho^{(1)}$.
1787: At this threshold the system undergoes a discontinuous transition
1788: from an imbalanced fluid phase,
1789: where both species have densities $\rho^{(1)}$ and $\rho^{(2)}_0$
1790: larger than the critical density, to an imbalanced condensate
1791: coexisting with a balanced critical fluid
1792: with densities $\rho^{(1)}=\rho^{(2)}=\rho_c$.
1793: 
1794: Finally, the observed rapid growth of the threshold density
1795: $\rho_0$ with the di\-men\-sio\-na\-li\-ty raises the question of the behavior
1796: of the asymmetric target process in high dimensions.
1797: At this point let us emphasize that the absence of a stationary-state product
1798: measure for the target process
1799: is a rather subtle effect which needs the conjunction of several ingredients,
1800: and chiefly the presence of a bias.
1801: This feature cannot be present in mean-field geometries
1802: such as the complete graph,
1803: so that the dynamical threshold behavior of the model in high dimensions
1804: is not expected to smoothly converge to a well-defined mean-field limit.
1805: 
1806: \section*{References}
1807: 
1808: \begin{thebibliography}{99}
1809: 
1810: \bibitem{loan}
1811: O'Loan O J, Evans M R and Cates M E, 1998 Phys. Rev. E {\bf 58} 1404
1812: 
1813: \bibitem{evans1}
1814: Evans M R, 2000 Braz. J. Phys. {\bf 30} 42
1815: 
1816: \bibitem{wis1}
1817: Kafri Y, Levine E, Mukamel D, Sch\"utz G M and T\"or\"ok J, 2002 Phys. Rev.
1818: Lett. {\bf 89} 035702
1819: 
1820: \bibitem{cg}
1821: Godr\`eche C, 2003 J. Phys. A {\bf 36} 6313
1822: 
1823: \bibitem{gross}
1824: Grosskinsky S, Sch\"utz G M and Spohn H, 2003 J. Stat. Phys. {\bf 113} 389
1825: 
1826: \bibitem{evans2}
1827: Evans M R and Hanney T, 2005 J. Phys. A {\bf 38} R195
1828: 
1829: \bibitem{glcond}
1830: Godr\`eche C and Luck J M, 2005 J. Phys. A {\bf 38} 7215
1831: 
1832: \bibitem{zeta1}
1833: Bialas P, Burda Z and Johnston D, 1997 Nucl. Phys. B {\bf 493} 505
1834: \nonum
1835: Bialas P, Burda Z and Johnston D, 1999 Nucl. Phys. B {\bf 542} 413
1836: \nonum
1837: Bialas P, Bogacz L, Burda Z and Johnston D, 2000 Nucl. Phys. B {\bf 575} 599
1838: 
1839: \bibitem{zeta2}
1840: Drouffe J M, Godr\`eche C and Camia F, 1998 J. Phys. A {\bf 31} L19
1841: 
1842: \bibitem{zeta3}
1843: Godr\`eche C and Luck J M, 2001 Eur. Phys. J. B {\bf 23} 473
1844: 
1845: \bibitem{barc}
1846: Godr\`eche C and Luck J M, 2002 J. Phys. Cond. Matt. {\bf 14} 1601
1847: 
1848: \bibitem{lux}
1849: Godr\`eche C, 2007 Lect. Notes Phys. {\bf 716} 216
1850: (Berlin: Springer) (cond-mat/0604276)
1851: 
1852: \bibitem{maj}
1853: Majumdar S N, Evans M R and Zia R K P, 2005 Phys. Rev. Lett. {\bf 94} 180601
1854: \nonum
1855: Evans M R, Majumdar S N and Zia R K P, 2006 J. Stat. Phys. {\bf 123} 357
1856: 
1857: \bibitem{spitz}
1858: Spitzer F, 1970 Advances in Math. {\bf 5} 246
1859: 
1860: \bibitem{andj}
1861: Andjel E D, 1982 Ann. Prob. {\bf 10} 525
1862: 
1863: \bibitem{kelly}
1864: Kelly F, 1979 {\it Reversibility and Stochastic Networks} (Chichester: Wiley)
1865: 
1866: \bibitem{cocozza}
1867: Cocozza-Thivent C, 1985 Z. Wahr. {\bf 70} 509
1868: 
1869: \bibitem{pairw}
1870: Sch\"utz G M, Ramaswamy R and Barma M, 1996 J. Phys. A {\bf 29} 837
1871: 
1872: \bibitem{kls3}
1873: Luck J M and Godr\`eche C, 2006 J. Stat. Mech. P08009
1874: 
1875: \bibitem{ehm}
1876: Evans M R, Hanney T and Majumdar S N, 2006 Phys. Rev. Lett. {\bf 97} 010602
1877: 
1878: \bibitem{zrp2}
1879: Godr\`eche C, 2006 J. Phys. A {\bf 39} 9055
1880: 
1881: \end{thebibliography}
1882: \end{document}
1883: