cond-mat0412077/1.tex
1: \documentclass[12pt]{iopart}
2: \usepackage{graphicx}
3: \begin{document}
4: \hyphenation{fer-ro-mag-net fer-ro-mag-nets meta-sta-ble meta-sta-bi-lity
5: con-fi-gu-ra-tion con-fi-gu-ra-tions expo-nen-tially}
6: 
7: \newcommand{\beq}{\begin{equation}}
8: \newcommand{\eeq}{\end{equation}}
9: \newcommand{\bea}{\begin{eqnarray}}
10: \newcommand{\eea}{\end{eqnarray}}
11: 
12: \newcommand{\abs}[1]{\vert#1\vert}
13: \newcommand{\ap}{{\rm ap}}
14: \newcommand{\bin}[2]{{#1\choose#2}}
15: \newcommand{\bl}{\circ{\hskip .21mm}}
16: \renewcommand{\c}{{\rm c}}
17: \newcommand{\ca}{{\rm S}}
18: \newcommand{\co}{{\rm co}}
19: \newcommand{\conf}{{\rm conf}}
20: \renewcommand{\d}{{\rm d}}
21: \newcommand{\del}{\delta}
22: \newcommand{\diag}{{\rm diag}}
23: \newcommand{\dmean}[1]{\langle#1\rangle}
24: \newcommand{\dpar}{\partial}
25: \newcommand{\dyn}{{\rm dyn}}
26: \renewcommand{\e}{{\rm e}}
27: \newcommand{\eps}{\varepsilon}
28: \newcommand{\eq}{{\rm eq}}
29: \newcommand{\frad}[2]{\displaystyle{\displaystyle#1\over\displaystyle#2}}
30: \newcommand{\g}{\gamma}
31: \newcommand{\he}{{\rm H}}
32: \renewcommand{\i}{{\rm i}}
33: \newcommand{\im}{{\mathop{\rm Im\ }}}
34: \newcommand{\lam}{\lambda}
35: \newcommand{\lra}{\Longleftrightarrow}
36: \renewcommand{\max}{{\rm max}}
37: \newcommand{\mean}[1]{\langle#1\rangle}
38: \newcommand{\no}{\bullet{\hskip .21mm}}
39: \newcommand{\prob}{{\rm Prob}}
40: \newcommand{\s}{\sigma}
41: \renewcommand{\sp}{{\rm sp}}
42: \renewcommand{\t}{\tau}
43: \newcommand{\var}{\mathop{\rm Var}}
44: \newcommand{\vecn}{{\bf n}}
45: \newcommand{\vecz}{{\bf 0}}
46: \newcommand{\w}{\omega}
47: \newcommand{\C}{{\cal C}}
48: \newcommand{\En}{{\cal E}}
49: \newcommand{\F}{{\cal F}}
50: \renewcommand{\H}{{\cal H}}
51: \newcommand{\I}{{\cal I}}
52: \newcommand{\HB}{{\rm HB}}
53: \newcommand{\M}{{\rm M}}
54: \newcommand{\N}{{\cal N}}
55: \renewcommand{\S}{\Sigma}
56: \newcommand{\T}{{\cal T}}
57: \eqnobysec
58: 
59: \title[Metastability in zero-temperature dynamics]
60: {Metastability in zero-temperature dynamics: Statistics of attractors}
61: \author{C Godr\`eche\dag\ and J M Luck\ddag}
62: 
63: \address{\dag\ Service de Physique de l'\'Etat Condens\'e,
64: CEA Saclay, 91191~Gif-sur-Yvette cedex, France}
65: 
66: \address{\ddag\ Service de Physique Th\'eorique\footnote{URA 2306 of CNRS},
67: CEA Saclay, 91191~Gif-sur-Yvette cedex, France}
68: 
69: \begin{abstract}
70: The zero-temperature dynamics of simple models such as Ising ferro\-magnets
71: provides, as an alternative to the mean-field situation,
72: interesting examples of dynamical systems with many attractors
73: (absorbing configurations, blocked confi\-gurations,
74: zero-temperature metastable states).
75: After a brief review of metastability in the mean-field ferromagnet
76: and of the droplet picture,
77: we focus our attention onto zero-temperature single-spin-flip dynamics
78: of ferromagnetic Ising models.
79: The situations leading to metastability are characterized.
80: The statistics and the spatial structure of the at\-trac\-tors thus obtained
81: are investigated, and put in perspective with uniform a priori ensembles.
82: We review the vast amount of exact results available in one dimension,
83: and present original results on the square and honeycomb lattices.
84: \end{abstract}
85: 
86: \pacs{05.70.Ln, 64.10.+h, 64.60.My}
87: \eads{\mailto{godreche@drecam.saclay.cea.fr},\mailto{luck@spht.saclay.cea.fr}}
88: \maketitle
89: 
90: \section{Introduction}
91: 
92: The non-equilibrium dynamics observed in a variety of systems,
93: ranging from glasses to granular media,
94: with its characteristic features of glassiness and aging~\cite{ange,glassy},
95: is often thought of as the motion of a particle in a complex energy landscape,
96: with many valleys and barriers~\cite{gold}.
97: This picture fully applies in the mean-field situation,
98: where barriers are infinitely high
99: and valleys appear as metastable states~\cite{tap,ks}.
100: For finite-dimensional systems with short-range interactions,
101: barrier heights and lifetimes are always finite at finite temperature,
102: so that metastability becomes a matter of time scales~\cite{gs,bir}.
103: 
104: The main goal of the present paper is to emphasize that the
105: zero-temperature dynamics of simple models such as Ising ferromagnets
106: provides, as an alternative to the mean-field situation,
107: interesting examples of dynamical systems with many attractors
108: (absorbing configurations, blocked configurations,
109: zero-temperature metastable states).
110: For completeness, we first give a brief review of the phenomenon
111: of metastability in the mean-field ferromagnet (Section~\ref{mft}),
112: and of the droplet picture implying that barrier heights become finite
113: for short-range interactions (Section~\ref{drop}).
114: We then turn to an overview of the configurational entropy
115: in disordered and complex systems,
116: and of the so-called Edwards hypothesis (Section~\ref{complex}).
117: The rest of this paper is devoted to zero-temperature single-spin-flip dynamics
118: of ferromagnetic Ising models.
119: We characterize the dynamics which lead to zero-temperature metastability,
120: and investigate the structure of the attractors thus obtained.
121: After a general presentation (Section~\ref{gene}),
122: we review the extensive amount of exact results available in one dimension
123: (\cite{ldt,prbr}, and especially~\cite{usrsa}) (Section~\ref{chain}),
124: and present original results on the square and honeycomb lattices
125: (Section~\ref{two}).
126: We end up with a brief discussion (Section~\ref{dis}).
127: 
128: \section{Metastability in the mean-field ferromagnet}
129: \label{mft}
130: 
131: The concepts of metastability and of spinodal line
132: date back to the early days of Thermodynamics,
133: with the seminal work of Gibbs~\cite{gibbs},
134: in the context of phase separation in simple systems such as fluids.
135: In order to summarize these early developments,
136: let us use the more modern language of the Landau theory
137: for a ferromagnet~\cite{ll}.
138: At the mean-field level,
139: i.e., neglecting any spatial dependence of the order parameter,
140: a ferromagnetic system of Ising type is
141: described by the Landau free energy density
142: \beq
143: F(M)=A(T-T_\c)\frac{M^2}{2}+B\frac{M^4}{4}-HM,
144: \label{gibbs}
145: \eeq
146: where $T$ is the temperature, $T_\c$ the Curie temperature,
147: $A$ and $B$ are positive phe\-no\-me\-no\-lo\-gi\-cal parameters,
148: and $H$ is the applied magnetic field.
149: The magnetization $M(T,H)$ is such that $F(M)$ is stationary.
150: It is therefore given by the equation of state
151: \beq
152: \frac{\dpar F}{\dpar M}=A(T-T_\c)M+BM^3-H=0.
153: \label{state}
154: \eeq
155: The corresponding magnetic susceptibility $\chi$ is such that
156: \beq
157: \frac{1}{\chi}=\frac{\dpar^2F}{\dpar M^2}=A(T-T_\c)+3BM^2.
158: \label{chi}
159: \eeq
160: 
161: At fixed temperature $T<T_\c$, the solutions to~(\ref{state})
162: are of the following three types:
163: 
164: \noindent (I) {\it stable}: absolute (global) minimum of the free energy,
165: for $\abs{M}>M_0(T)$,
166: 
167: \noindent (II) {\it metastable}: relative (local) minimum of the free energy,
168: for $M_\sp(T)<\abs{M}<M_0(T)$,
169: 
170: \noindent (III) {\it unstable}: maximum of the free energy,
171: for $\abs{M}<M_\sp(T)$.
172: 
173: \begin{figure}[htb]
174: \begin{center}
175: \includegraphics[angle=90,width=.5\linewidth]{figmh.eps}
176: \vskip 5mm
177: \includegraphics[angle=90,width=.5\linewidth]{figmt.eps}
178: \caption{\small
179: The three types of solutions to the equation of state~(\ref{state}).
180: Upper panel: Plot of the magnetization $M(H,T)$ against the magnetic field $H$,
181: for a fixed temperature $T<T_\c$.
182: Thin full line: stable solution (I).
183: Thick full line: metastable solution (II).
184: Thick dotted line: unstable solution (III).
185: Lower panel: Phase diagram in the $M$--$T$ plane.
186: Phases I to III correspond to the three types of solutions.
187: Full line: spontaneous magnetizations $M=\pm M_0(T)$.
188: Dashed line: spinodal magnetizations $M=\pm M_\sp(T)$.}
189: \label{figm}
190: \end{center}
191: \end{figure}
192: 
193: Figure~\ref{figm} illustrates this discussion.
194: When the magnetic field $H$ is positive,
195: the stable solution $M(H,T)$ describes the magnetization of the $+$ phase.
196: If $H$ decreases to zero,
197: the magnetization takes its spontaneous value $M(0,T)=M_0(T)$.
198: Coexistence is reached: the $+$ phase becomes degenerate with the $-$ phase.
199: If $H$ decreases further to
200: a small negative value, the $+$ phase still exists but becomes metastable.
201: Its magnetization $M(H,T)$ can be analytically continued
202: all the way until the endpoint of the metastability region,
203: which is reached for $\dpar F/\dpar M=\dpar^2F/\dpar M^2=0$.
204: The corresponding values of $M$ and $H$ are respectively called the spinodal
205: magnetization $M_\sp(T)$ and the coercive magnetic field $H_\co(T)$.
206: All physical quantities are singular at the spinodal endpoint.
207: The spontaneous magnetization, spinodal magnetization,
208: and coercive field read
209: \bea
210: M_0(T)=M(0,T)=\left(\frac{A(T_\c-T)}{B}\right)^{1/2},\nonumber\\
211: M_\sp(T)=M(-H_\co(T),T)
212: =\left(\frac{A(T_\c-T)}{3B}\right)^{1/2}=\frac{M_0(T)}{\sqrt{3}},\\
213: H_\co(T)=\left(\frac{4A^3(T_\c-T)^3}{27B}\right)^{1/2}
214: =\frac{2BM_0(T)^3}{3\sqrt{3}}.\nonumber
215: \eea
216: 
217: The metastable solution
218: is separated from the stable one by an extensive free energy barrier,
219: whose height $\Omega\,\Delta F$ grows proportionally to the volume
220: $\Omega$ of the system.
221: The activation energy density $\Delta F$ is the difference in $F(M,T)$
222: between the metastable solution (II) and the unstable one (III).
223: It has a finite limit
224: \beq
225: \Delta F=\frac{A^2(T_\c-T)^2}{4B}
226: \eeq
227: at coexistence, whereas it vanishes at the spinodal endpoint.
228: 
229: The continuation of the stable phase to the metastable one
230: corresponds to an analytical continuation in the mathematical sense.
231: The magnetization $M(H,T)$, defined as the stable solution to~(\ref{state})
232: for $H>0$, is indeed analytic at $H=0$.
233: The power-series expansion
234: \beq
235: M(H,T)=\sum_{n\ge0}a_n(T)\,H^n
236: \label{aims}
237: \eeq
238: has a finite radius of convergence, equal to $\abs{H_\co(T)}$.
239: 
240: \section{Finite-dimensional ferromagnet: Metastability vs.~droplet picture}
241: \label{drop}
242: 
243: It has been realized in the 1960s that metastability
244: is an artefact of the mean-field approximation.
245: For realistic Ising ferromagnetic systems with short-range interactions
246: in a finite-dimensional space,
247: Fisher~\cite{fisher} and Langer~\cite{langer}
248: have shown that the magnetization $M(H,T)$ of the $+$ phase,
249: or, equivalently, its free energy $F(H,T)$, is {\it not} analytic
250: at the coexistence point $H=0$.
251: Their argument relies on the droplet theory initiated by Frenkel~\cite{frenkel}.
252: Consider, in a $d$--dimensional ferromagnetic model below its Curie temperature,
253: a spherical droplet with radius $R$
254: of the $-$ phase ($M=-M_0(T)$) inside an infinite $+$ phase
255: ($M=+M_0(T)$).
256: The corresponding excess free energy
257: \beq
258: f(R)=2\w R^dHM_0(T)+d\w R^{d-1}\s_0(T)
259: \eeq
260: is the sum of a volume term, proportional to the volume $\w R^d$ of the droplet,
261: to the spontaneous magnetization $M_0(T)$, and to the magnetic field $H$,
262: and of a surface term, proportional to the surface area $d\w R^{d-1}$
263: of the droplet and to the interfacial free energy $\s_0(T)$.
264: Here $\w=2\pi^{d/2}/\Gamma(d/2)$ denotes the volume of the
265: $d$--dimensional unit sphere.
266: If the magnetic field $H$ is positive,
267: the excess free energy is a positive and growing function of the radius $R$.
268: If the magnetic field is negative, however,
269: the excess free energy reaches a maximum~\cite{frenkel},
270: \beq
271: f_\c=f(R_\c)=\w\s_0(T)^d\left(\frac{d-1}{2\abs{H}M_0(T)}\right)^{d-1},
272: \eeq
273: for a finite critical radius
274: \beq
275: R_\c=\frac{(d-1)\s_0(T)}{2\abs{H}M_0(T)}.
276: \eeq
277: 
278: The existence of a critical droplet with a {\it finite} radius $R_\c$,
279: and hence of a {\it finite} barrier height $f_\c$,
280: has far-reaching consequences, both in statics and in dynamics.
281: The reduced barrier height $f_\c/T$ of the critical droplet
282: (in units where Boltzmann's constant is unity)
283: plays a role similar to that of the reduced instanton action
284: $S/\hbar$ in Quantum Mechanics~\cite{zinn}.
285: 
286: \begin{itemize}
287: 
288: \item {\it Statics.}
289: The free energy $F(H)$ of the $+$ phase is not analytic at $H=0$.
290: In other words, the formal power-series expansion~(\ref{aims}) is divergent:
291: its radius of convergence vanishes.
292: It can indeed be shown that $F(H)$ has a branch cut starting at $H=0$,
293: with an exponentially small imaginary part for $H<0$, of the form
294: \beq
295: \im F(H+\i0)\sim\exp\left(-\frac{f_\c}{T}\right).
296: \label{imf}
297: \eeq
298: 
299: \item {\it Dynamics.}
300: If the magnetic field is instantaneously turned from a positive
301: to a small negative value, the `metastable' $+$ phase
302: has a finite lifetime $\tau$,
303: whose order of magnitude is the inverse nucleation rate of a critical droplet,
304: given by an Arrhenius law in terms of the above barrier height:
305: \beq
306: \frac{1}{\tau}\sim\exp\left(-\frac{f_\c}{T}\right).
307: \label{tauinv}
308: \eeq
309: 
310: \end{itemize}
311: 
312: Near coexistence ($\abs{H}\to0$), the estimates~(\ref{imf})
313: and~(\ref{tauinv}) have an
314: exponentially small essential singularity of the form
315: \beq
316: \frac{1}{\tau}\sim\im F(H+\i0)\sim\exp\left(-\frac{\w\s_0(T)^d}{T}
317: \left(\frac{d-1}{2\abs{H}M_0(T)}\right)^{d-1}\right).
318: \eeq
319: 
320: To sum up, for ferromagnets and similar systems,
321: the metastability scenario {\it stricto sensu}
322: (analytic free energy at the coexistence point,
323: infinite lifetime in the metastable region, sharp spinodal line)
324: is a peculiar feature of the mean-field and of the zero-temperature situations.
325: For systems with short-range interactions,
326: at finite temperature in finite dimension,
327: the barrier height $f_\c$ is finite all over the would-be metastable region.
328: The free energy has an essential singularity at the coexistence point,
329: and the lifetime of the metastable phase is finite,
330: although it becomes expo\-nen\-tially large near coexistence.
331: The spinodal line therefore becomes a crossover between
332: the fast decay of an unstable phase (no barrier height)
333: and the slow decay of an approximately metastable phase (finite barrier height).
334: This crossover phenomenon has been recently investigated
335: by means of a sophisticated numerical approach~\cite{bustillos}.
336: 
337: \section{Metastable states and complexity in disordered and complex systems}
338: \label{complex}
339: 
340: From the 1970s on, a variety of complex systems,
341: such as structural glasses, spin glasses, and granular materials,
342: have been shown to possess many metastable states at low enough temperature,
343: or high enough density.
344: The number $\N(N;E)$ of metastable states with given energy density $E$
345: typically grows exponentially with the system size, as
346: \beq
347: \N(N;E)\sim\exp(N\,S_\conf(E)).
348: \label{nconf}
349: \eeq
350: The quantity $S_\conf(E)$ is referred to
351: as the configurational entropy, or complexity~\cite{sconf}.
352: 
353: Most investigations of the complexity
354: were motivated by dynamical considerations.
355: The dynamics of the above systems
356: in their low-temperature or high-density regime
357: often turns out to be so slow
358: that the system falls out of equilibrium~\cite{ange},
359: and exhibits aging and other characteristic features
360: of glassy dynamics~\cite{glassy}.
361: It has been proposed long ago to describe slow relaxational dynamics
362: in terms of the motion of a particle
363: in a complex energy (or free energy) landscape~\cite{gold},
364: with many valleys separated by barriers.
365: Several approaches have been developed,
366: in order to make this heuristic picture more precise,
367: and chiefly to identify these valleys.
368: Metastable states have thus been rediscovered under various names
369: and with various definitions, in different contexts:
370: valleys~\cite{ks}, pure states~\cite{tap,ktw},
371: inherent structures~\cite{sw}, quasi-states~\cite{fv}.
372: 
373: In the mean-field situation,
374: all these metastable states are separated by extensive barriers,
375: so that they have an infinite lifetime,
376: just as the unfavored phase in an Ising ferromagnet.
377: For finite-dimensional systems with short-range interactions,
378: barrier heights and lifetimes are always finite at finite temperature,
379: so that metastability becomes a matter of time scales.
380: Furthermore, from a dynamical viewpoint, the various concepts
381: of metastability recalled above are not equivalent~\cite{gs,bir}.
382: 
383: Whenever metastable states live forever,
384: i.e., either in the mean-field situation or at zero temperature,
385: a natural question concerns their {\it dynamical weights}.
386: Consider a system instantaneously quenched into its low-temperature
387: or high-density regime, from a randomly chosen initial configuration.
388: {\it Does the system sample all the possible metastable states
389: with given energy density with equal statistical weights,
390: i.e., with a uniform or flat measure, or, to the contrary,
391: does any detailed feature, such as the shape or size of the attraction basin
392: of each metastable state, matter?}
393: This question is of interest for any dynamical system having a large number
394: of attractors.
395: 
396: A similar question has been addressed in many works on the tapping
397: of granular systems~\cite{ldt,bm,pbt}.
398: Under tapping, a granular material constantly jumps
399: from a blocked configuration to a nearby one.
400: In this context, Edwards~\cite{edwards} proposed to describe situations
401: such as slow compaction dynamics,
402: or the steady-state dynamics under gentle tapping,
403: by means of a flat ensemble average over the {\it a priori ensemble}
404: of all the blocked configurations of the grains with prescribed density.
405: 
406: Extending the range of application of this idea far beyond its original scope,
407: the so-called {\it Edwards hypothesis} commonly refers to the assumption
408: that all the metastable states with given energy density are equivalent.
409: This hypothesis has two consequences.
410: First, the value of an observable can be obtained by a flat average
411: over the a priori ensemble, or Edwards ensemble, of all those metastable states.
412: Second, the configurational temperature, or Edwards temperature,
413: $T_\conf=(\d S_\conf/\d E)^{-1}$,
414: is expected to share the usual thermodynamical properties of a temperature,
415: and especially to coincide with the effective temperature
416: involved in the generalized fluctuation-dissipation formula
417: in the appropriate temporal regime~\cite{fdt}.
418: The concept of ergodicity, and the resulting thermodynamical construction,
419: therefore hold, as in equilibrium situations,
420: up to the replacement of configurations by metastable states,
421: and of temperature by the configurational temperature~$T_\conf$.
422: 
423: The Edwards hypothesis has been shown to be valid
424: for the slow relaxational dynamics of some mean-field models~\cite{fv},
425: where metastable states are indeed explored with a flat measure.
426: This hypothesis has also been found to hold for finite-dimensional systems,
427: at least as a good numerical approximation,
428: both in the steady-state of tapped systems
429: in the regime of weak tapping intensities~\cite{ldt,bm,pbt},
430: and in the regime of slow relaxational dynamics of various models~\cite{ba}.
431: 
432: Besides the mean-field geometry,
433: another physical situation where metastable states
434: are unambiguously defined is the zero-temperature limit,
435: where no barrier can be crossed at all.
436: The metastable states are then defined as
437: the blocked configurations under the chosen dynamics.
438: For instance, for an Ising model with single-spin-flip dynamics,
439: a metastable state is a configuration where each spin is aligned
440: with its local field, whenever the latter is non-zero.
441: The rest of this paper is devoted to the statistical analysis
442: of the blocked configurations thus generated,
443: before we come back to the so-called Edwards hypothesis in Section~\ref{dis}.
444: 
445: \section{Zero-temperature single-spin-flip dynamics of the Ising model:
446: Generalities}
447: \label{gene}
448: 
449: This section is devoted to an overview of zero-temperature
450: dynamics of ferromagnetic Ising models.
451: In the two following sections, we review the extensive available results
452: in one dimension (\cite{ldt,prbr}, and especially~\cite{usrsa}),
453: and present novel results in two dimensions.
454: 
455: The ferromagnetic Ising model is defined by the Hamiltonian
456: \beq
457: \H=-\sum_{(m,n)}\s_m\s_n,
458: \label{hamf}
459: \eeq
460: where the $\s_n=\pm1$ are classical Ising spins
461: sitting at the vertices of a regular lattice with coordination number~$z$,
462: and the sum runs over all pairs $(m,n)$ of nearest neighbors.
463: 
464: Consider single-spin-flip dynamics (i.e., Glauber dynamics, in a broad sense)
465: in continuous time.
466: Each spin is flipped ($\s_n\to-\s_n)$ with a rate $W(\del\H)$ per unit time.
467: This rate is assumed to only depend on the energy difference
468: implied in the flip,
469: \beq
470: \del\H=2h_n\s_n,
471: \eeq
472: where
473: \beq
474: h_n=-\frac{\dpar\H}{\dpar\s_n}=\sum_{m(n)}\s_m,
475: \eeq
476: is the local field acting on spin $n$.
477: The sum in the above definition runs over the $z$ nearest neighbors $m$
478: of site~$n$.
479: 
480: The condition of detailed balance at temperature $T=1/\beta$
481: with respect to the Hamiltonian $\H$ reads
482: \beq
483: \frac{W(\del\H)}{W(-\del\H)}=\exp(-\beta\del\H).
484: \label{deba}
485: \eeq
486: The two most usual choices of flipping rates,
487: or flipping probabilities in the case of discrete updates,
488: \beq
489: \matrix{
490: \hbox{Heat-bath:}\hfill& W_\HB(\del\H)=\frad{1}{1+\exp(2\beta\del\H)},
491: \hfill\cr
492: \hbox{Metropolis:}\quad\hfill& W_\M(\del\H)=\min(1,\exp(-\beta\del\H)),\hfill}
493: \label{hbm}
494: \eeq
495: obey the detailed balance condition~(\ref{deba}) at any temperature.
496: It is, however, less well-known that there is a vast family of dynamical rates,
497: besides these two choices, which obeys the condition~(\ref{deba}).
498: Indeed, the energy difference $\del\H$ may assume $z+1$
499: equidistant integer values:
500: \beq
501: \del\H=2z,\ 2z-4,\ \dots,\ -2z,
502: \label{list}
503: \eeq
504: so that there are a priori $z+1$ different rates.
505: If the coordination number $z$ is even (resp.~odd),
506: the condition~(\ref{deba}) only gives $z/2$ (resp.~$(z+1)/2$)
507: equations for the $z+1$ unknown rates,
508: so that there remain $(z+2)/2$ (resp.~$(z+1)/2$) free parameters
509: (one of them corresponding to the absolute time scale).
510: 
511: The arbitrariness described above persists in the zero-temperature limit.
512: The condition of detailed balance~(\ref{deba}) indeed reads
513: \beq
514: W(\del\H)=0\quad\hbox{for all }\del\H>0,
515: \label{baz}
516: \eeq
517: whereas nothing is imposed on the rates with $\del\H\le0$.
518: For the sake of definiteness,
519: we focus our attention onto the class of zero-temperature dynamics defined
520: by the following flipping rates:
521: \beq
522: W(\del\H)=\left\{\matrix{
523: 0\hfill&\hbox{if }\del\H>0,\cr
524: W_0\quad\hfill&\hbox{if }\del\H=0,\cr
525: 1\hfill&\hbox{if }\del\H<0.}\right.%}
526: \label{wz}
527: \eeq
528: The rate~$W_0$ relative to the {\it free} spins is kept as a parameter.
529: A free spin is defined as a spin $\s_n$
530: which is subjected to a zero local field ($h_n=0$, hence $\del\H=0$).
531: Free spins can only exist if the coordination number $z$ is even,
532: so that 0 belongs to the list~(\ref{list}).
533: The zero-temperature limits of the heat-bath and Metropolis rates~(\ref{hbm})
534: are respectively ${W_0}_\HB=1/2$ and ${W_0}_\M=1$.
535: 
536: The zero-temperature dynamics of ferromagnetic Ising systems
537: is far from being a trivial problem in general~\cite{red,ns}.
538: The single-spin-flip dynamics defined by~(\ref{wz}) is a descent dynamics
539: (i.e., every move strictly lowers the total energy)
540: in the following two situations:
541: 
542: \begin{itemize}
543: 
544: \item $z$ is odd, so that there are no free spins at all.
545: \item $z$ is even and $W_0=0$.
546: The corresponding dynamics is said to be {\it constrained}.
547: As such it belongs to the class of
548: {\it kinetically constrained} systems~\cite{barc,riso}.
549: 
550: \end{itemize}
551: 
552: In both situations, every spin flips a finite number of times,
553: and the system gets trapped in a non-trivial attractor
554: (absorbing configuration, blocked confi\-guration,
555: zero-temperature metastable state).
556: 
557: \section{Constrained zero-temperature single-spin-flip dynamics
558: on the Ising chain}
559: \label{chain}
560: 
561: The one-dimensional situation of a chain of spins corresponds to the
562: smallest possible value of the coordination number ($z=2$),
563: so that the list of values of the energy difference is 4, 0, and $-4$.
564: The heat-bath dynamics has been solved exactly by Glauber~\cite{glau}.
565: 
566: Let us focus our attention onto zero-temperature dynamics.
567: For any non-zero value of the rate $W_0$ corresponding to free spins,
568: the dynamics belongs to the universality class of the zero-temperature
569: Glauber model.
570: This is a prototypical example of phase ordering by domain growth
571: (coarsening)~\cite{bray}.
572: The typical size of ordered domains of consecutive $+$ and $-$ spins
573: grows as $L(t)\sim t^{1/2}$.
574: The energy density $E(t)=-1+2/L(t)$ therefore
575: relaxes to its ground-state value as $t^{-1/2}$.
576: The particular value $W_0=0$ corresponds to
577: the constrained zero-temperature Glauber dynamics,
578: which has been investigated in~\cite{ldt,prbr}, and especially in~\cite{usrsa}.
579: Hereafter we summarize the main results of these investigations.
580: 
581: \subsection{Mapping onto the dimer RSA model}
582: 
583: In the constrained zero-temperature Glauber dynamics,
584: the only possible moves are flips of isolated spins:
585: \beq
586: -+-\;\to\;---,\quad+-+\;\to\;+++.
587: \label{gls}
588: \eeq
589: Each move suppresses two consecutive unsatisfied bonds.
590: The system therefore eventually reaches a blocked configuration,
591: where there is no isolated spin, i.e., up and down spins form clusters
592: whose length is at least two.
593: Equivalently, each unsatisfied bond (or domain wall) is isolated.
594: These blocked configurations are the zero-temperature analogues
595: of metastable states.
596: 
597: The dynamics can be recast in the following illuminating way.
598: Going to the dual lattice, where dual sites represent bonds,
599: let us represent unsatisfied bonds as empty dual sites,
600: and satisfied bonds as occupied dual sites:
601: \beq
602: \left\{\matrix{
603: \t_n=\s_n\s_{n+1}=-1\lra\bl,\hfill\cr
604: \t_n=\s_n\s_{n+1}=+1\lra\no,\hfill
605: }\right.%}
606: \label{corrbond}
607: \eeq
608: so that the moves~(\ref{gls}) read
609: \beq
610: \bl\bl\to\no\no.
611: \label{rsadim}
612: \eeq
613: This mapping shows at once that the dynamics is fully irreversible,
614: in the sense that each spin flips at most once
615: during the whole history of the sample.
616: 
617: The dynamics~(\ref{rsadim}) is identical to that of the
618: random sequential adsorption (RSA) of lattice dimers,
619: which has been considered long ago~\cite{flory,core}.
620: This is one of the simplest examples of RSA problems,
621: for which powerful analytical techniques are available in one dimension,
622: both on the lattice and in the continuum~\cite{rsa}.
623: A blocked configuration thus appears as a jammed state of the dimer RSA model,
624: as illustrated on the following example:
625: \bea
626: &&+---++--++++---+++-----+++--\cr
627: &&{\hskip 2.5mm}
628: \bl\no\no\bl\no\bl\no\bl\no\no\no\bl\no\no\bl
629: \no\no\bl\no\no\no\no\bl\no\no\bl\no
630: \eea
631: 
632: \subsection{Relaxation of the mean energy}
633: 
634: We consider as initial state the equilibrium state at finite temperature $T_0$.
635: Each bond variable is independently drawn from the binary distribution
636: \beq
637: \t_n=\left\{\matrix{
638: -1\hfill&(\bl)\hfill&\hbox{with probability}\;p,\hfill\cr
639: +1\hfill&(\no)\hfill&\hbox{with probability}\;1-p,\hfill
640: }\right.%}
641: \label{epsf}
642: \eeq
643: where the parameter $p$ is related to the energy density $E_0$
644: of the initial state and to the corresponding temperature $T_0$ by
645: \beq
646: \beta_0=\frac{1}{T_0}=\frac{1}{2}\,\ln\frac{1-p}{p},\quad E_0=-1+2p.
647: \eeq
648: 
649: It is a common feature of one-dimensional RSA problems~\cite{rsa}
650: that the densities of certain patterns obey closed rate equations.
651: In the present case,
652: the densities $p_\ell(t)$ of clusters made of exactly~$\ell$
653: consecutive unsatisfied bonds (empty dual sites)
654: obey the linear equations~\cite{usrsa,flory}
655: \beq
656: \frac{\d p_\ell(t)}{\d t}=-(\ell-1)p_\ell(t)+2\sum_{k\ge\ell+2}p_k(t)
657: \label{dpl}
658: \eeq
659: for $\ell\ge1$, with $p_\ell(0)=(1-p)^2p^\ell$.
660: These equations can be solved by making the Ansatz
661: $p_\ell(t)=a(t)\,z(t)^\ell$ for $\ell\ge1$.
662: The solution thus obtained,
663: \beq
664: p_\ell(t)=(1-p\e^{-t})^2\exp(2p(\e^{-t}-1))\,p^\ell\e^{-(\ell-1)t},
665: \label{plt}
666: \eeq
667: leads to
668: \beq
669: E(t)=-1+2\sum_{\ell\ge1}\ell\,p_\ell(t)=-1+2p\exp(2p(\e^{-t}-1)).
670: \eeq
671: Only clusters of length $\ell=1$ survive in the blocked configurations,
672: as could be expected.
673: These defects occur with a finite density,
674: \beq
675: p_1(\infty)=p\e^{-2p},
676: \eeq
677: so that the energy density of blocked configurations
678: takes a non-trivial value, which depends continuously
679: on the initial temperature via the parameter $p$~\cite{ldt,prbr,usrsa}:
680: \beq
681: E_\infty=E(\infty)=-1+2p\e^{-2p}.
682: \label{einf}
683: \eeq
684: This non-trivial dependence is a clear evidence
685: that the dynamics is not ergodic.
686: For an initial state close to the ferromagnetic ground-state
687: ($E_0\to-1$, i.e., $p\to0$), the behavior $E_\infty\approx E_0-4p^2$
688: is easily explained in terms of the clusters made of two empty sites.
689: The energy of blocked states then increases monotonically with $p$,
690: up to the maximum value $E_\infty=-1+\e^{-1}\approx-0.632121$,
691: for an uncorrelated (infinite-temperature)
692: initial state ($p=1/2$, i.e., $E_0=0$),
693: and then decreases monotonically with $p$, down to the value
694: $E_\infty=-1+2\e^{-2}\approx-0.729329$,
695: corresponding to an antiferromagnetically ordered initial state ($p=1$).
696: 
697: \subsection{Distribution of the blocking time}
698: 
699: The late stages of the dynamics are dominated
700: by an exponentially small density of surviving clusters made of two empty sites.
701: More precisely,~(\ref{plt}) shows that their density reads
702: $p_2(t)\approx\alpha\e^{-t}$, with $\alpha=p^2\e^{-2p}$.
703: The dynamics can therefore be effectively described by
704: a collection of $\alpha N$ such clusters,
705: each cluster decaying exponentially with unit rate, when a down spin flips.
706: The blocking time $T_N$ is the largest of the decay times of those clusters.
707: For a large sample, it is therefore distributed according to extreme-value
708: statistics~\cite{gumbel}:
709: \beq
710: T_N=\ln(\alpha N)+X_N,
711: \label{tx}
712: \eeq
713: where the fluctuation $X_N$ remains of order unity,
714: and is asymptotically distributed according to the Gumbel law:
715: $f(X)=\exp(-X-\e^{-X})$.
716: 
717: \subsection{Distribution of the final energy: Dynamics vs.~a priori ensemble}
718: 
719: The attractors of the zero-temperature constrained
720: single-spin-flip dynamics of the Ising chain are the spin configurations
721: where every unsatisfied bond is isolated.
722: The most natural statistical description of these attractors
723: is provided by the a priori ensemble
724: where all the blocked configurations with given energy density $E$
725: are taken with equal weights.
726: 
727: In this section, the exact distribution of the final energy
728: of the blocked con\-fi\-gu\-ra\-tions is compared to the prediction
729: of the a priori approach.
730: It turns out that there are two a priori predictions,
731: the second (refined) one being a far better approximation
732: than the first (naive) one.
733: 
734: For a finite chain made of $N$ spins,
735: the number of blocked configurations with exactly $M$ unsatisfied bonds,
736: i.e., empty dual sites, reads
737: \beq
738: \N_{N,M}=\bin{N-M+b}{M},
739: \label{bino}
740: \eeq
741: where $b=0$ (resp.~1) for a closed (resp.~open) chain.
742: This is indeed the number of ways of inserting the $M$ empty sites
743: in the $N-M+b$ spaces made available by the presence
744: of $N-M$ occupied sites, with at most one empty site
745: per available space~\cite{usrsa}.
746: In the limit of a large system ($M,N\to\infty$,
747: with a fixed ratio $M/N=(E+1)/2$), this number grows exponentially,
748: irrespective of $b$, as
749: \beq
750: \N(N;E)\sim\exp(N\,S_\ap(E)),
751: \eeq
752: in accord with~(\ref{nconf}).
753: The a priori configurational entropy reads~\cite{ldt,cri,bfs}
754: \beq
755: S_\ap(E)=E\ln(-2E)+\frac{1-E}{2}\ln(1-E)-\frac{1+E}{2}\ln(1+E).
756: \label{sap}
757: \eeq
758: 
759: This result can be alternatively derived
760: by means of the transfer-matrix method~\cite{baxter,tm}.
761: For a finite chain of $N$ spins, we introduce the partition function
762: \beq
763: Z_N=\sum_\C\e^{-\g\H(\C)},
764: \eeq
765: where the sum runs over all the blocked configurations $\C$,
766: as well as the partial partition functions $Z_N^\no$, $Z_N^\bl$
767: labeled by the prescribed value of the last bond, according to~(\ref{corrbond}).
768: The latter quantities obey the recursion
769: \beq
770: \pmatrix{Z_{N+1}^\no\cr Z_{N+1}^\bl}=\T\pmatrix{Z_N^\no\cr Z_N^\bl},\quad
771: \T=\pmatrix{\e^\g&\e^\g\cr\e^{-\g}&0}.
772: \eeq
773: The $2\times2$ transfer matrix $\T$ has eigenvalues
774: \beq
775: \lam_\pm=\frac{\e^\g\pm\sqrt{4+\e^{2\g}}}{2},
776: \eeq
777: so that $Z_N^\no\sim Z_N^\bl\sim\exp(N\ln\lam_+)$.
778: The configurational entropy $S_\ap(E)$
779: is therefore given by a Legendre transform:
780: \beq
781: \ln\lam_+(\g)+\g E-S_\ap(E)=0,
782: \quad E=-\frac{\d\ln\lam_+}{\d\g},
783: \quad\g=\frac{\d S_\ap}{\d E},
784: \label{thertrans}
785: \eeq
786: hence
787: \beq
788: E=-\frac{\e^\g}{\sqrt{4+\e^{2\g}}},\quad\g=\ln\frac{-2E}{\sqrt{1-E^2}},
789: \label{egge}
790: \eeq
791: so that~(\ref{sap}) is recovered.
792: 
793: The most naive prediction for the probability distribution $f(E)$
794: of the final energy density
795: is that $f(E)$ is proportional to the number $\N(N;E)$.
796: We thus obtain a large-deviation estimate of the form
797: \beq
798: f(E)\sim\exp(-N\,\S_\ap(E)),
799: \eeq
800: where the naive prediction for the large-deviation function $\S_\ap(E)$ reads
801: \beq
802: \S_\ap(E)=S_\max-S_\ap(E),
803: \label{signai}
804: \eeq
805: with
806: \beq
807: S_\max=\ln\frac{\sqrt{5}+1}{2}\approx0.481212
808: \label{smax}
809: \eeq
810: being the maximum of the configurational entropy $S_\ap(E)$.
811: This maximum is reached for
812: \beq
813: E_\max=-\frac{1}{\sqrt{5}}\approx-0.447214,
814: \label{eap}
815: \eeq
816: which is therefore the typical a priori energy density
817: of a blocked configuration.
818: 
819: A more refined prediction for the distribution of the final energy density,
820: introduced in the context of Kawasaki dynamics~\cite{uskawa},
821: resides on the following observation.
822: The mean energy density of blocked configurations,
823: as they are generated by the zero-temperature dynamics,
824: is equal to $E_\infty$, given by~(\ref{einf}),
825: which is in general different from the a priori value~(\ref{eap}).
826: Within the a priori formalism under consideration,
827: this difference is taken into account
828: by attributing to every configuration $\C$
829: an extra weight of the form $\exp(-\g_{\rm eff}NE(\C))$.
830: The effective inverse temperature $\g_{\rm eff}$ is chosen so that
831: the mean energy density
832: coincides with~(\ref{einf}), hence $\g_{\rm eff}=\g(E_\infty)$,
833: where $\g(E)=\d S_\ap/\d E$ is given in~(\ref{egge}).
834: This procedure amounts to replacing the configurational entropy $S_\ap(E)$,
835: which is maximal at $E=E_\max$, by the relative entropy
836: \beq
837: S_\ap(E\vert E_\infty)=S_\ap(E)-(E-E_\infty)\g(E_\infty),
838: \eeq
839: which is maximal at $E=E_\infty$ by construction.
840: The resulting prediction for the distribution
841: of the final energy density now involves
842: the refined large-deviation function
843: \beq
844: \S_\ap(E\vert E_\infty)=S_\ap(E_\infty)-S_\ap(E)+(E-E_\infty)\g(E_\infty).
845: \label{sigref}
846: \eeq
847: 
848: It turns out that the distribution of the final energy density
849: of the blocked configurations,
850: as they are generated by the zero-temperature dynamics,
851: has been evaluated exactly by analytical means~\cite{usrsa}.
852: We prefer to skip every detail of this rather lengthy derivation
853: and to state the result.
854: The distribution of the final energy density
855: is given by an exponential estimate of the form
856: \beq
857: f(E)\sim\exp(-N\,\S_\dyn(E)),
858: \eeq
859: where the large-deviation function $\S_\dyn(E)$
860: depends on the parameter $p$ characterizing the initial state,
861: and reads, in parametric form:
862: \bea
863: &&{\hskip -4.75mm}
864: \S_\dyn=\ln z+\frac{(1+(2p-1)z)^2-(z-1)^2\e^{4p z}}
865: {4p z^2\e^{2p z}(2(1-p)+(2p-1)z)}
866: \ln\frac{1+(2p-1)z+(1-z)\e^{2p z}}{1+(2p-1)z-(1-z)\e^{2p z}},\nonumber\\
867: &&
868: E=-1+\frac{(1+(2p-1)z)^2-(z-1)^2\e^{4p z}}
869: {2p z^2\e^{2p z}(2(1-p)+(2p-1)z)}.
870: \label{sigf}
871: \eea
872: 
873: \begin{figure}[htb]
874: \begin{center}
875: \includegraphics[angle=90,width=.5\linewidth]{figent.eps}
876: \caption{\small
877: Full line: plot of the dynamical entropy
878: of the ferromagnetic chain with constrained Glauber dynamics,
879: given by $\S_\dyn(E)$~(\ref{sigf}), against energy $E$,
880: for an uncorrelated initial state ($p=1/2)$.
881: Dash-dotted line: prediction~(\ref{signai}) of the naive a priori approach.
882: Dashed line: prediction~(\ref{sigref}) of the refined a priori approach
883: (after~\cite{usrsa}).}
884: \label{figent}
885: \end{center}
886: \end{figure}
887: 
888: Figure~\ref{figent} shows a comparison between the dynamical entropy
889: $\S_\dyn(E)$, given by~(\ref{sigf}),
890: for an uncorrelated (infinite-temperature) initial state ($p=1/2)$,
891: and the predictions~(\ref{signai}) of the naive a priori approach
892: and~(\ref{sigref}) of refined a priori approach.
893: The refined prediction turns out to be a good approximation
894: to the dynamical entropy, although it is not exact.
895: The following particular values allow a quantitative comparison.
896: At the ground-state energy density $E=-1$, one has
897: \beq
898: \S_\dyn(-1)=\ln z_\c(p),\quad
899: \S_\ap(-1\vert E_\infty)=\ln\frac{1-p\e^{-2p}}{1-2p\e^{-2p}}
900: \eeq
901: with
902: \beq
903: 1+(2p-1)z_\c+(1-z_\c)\e^{2p z_\c}=0,
904: \label{zcp}
905: \eeq
906: i.e., for $p=1/2$,
907: $\S_\dyn(-1)\approx0.245660$, $\S_\ap(-1\vert E_\infty)\approx0.255408$.
908: At the maximum allowed energy $E=0$, one has
909: \beq
910: \S_\dyn(0)=-\frac{1}{2}\ln p(1-p),\quad
911: \S_\ap(0\vert E_\infty)=\frac{1}{2}\ln\frac{1-p\e^{-2p}}{p\e^{-2p}},
912: \eeq
913: i.e., for $p=1/2$,
914: $\S_\dyn(0)\approx0.693147$, $\S_\ap(0\vert E_\infty)\approx0.744940$.
915: 
916: Another quantity of interest is the dynamical specific heat $C_\dyn$,
917: which is defined as the curvature of the dynamical entropy
918: at the mean energy density,
919: \beq
920: \S_\dyn(E)\approx\frac{(E-E_\infty)^2}{2C_\dyn},
921: \eeq
922: so that the energy variance (in the sense of sample to sample fluctuations
923: of the observed energy per spin in the blocked configurations) behaves as
924: \beq
925: \var E\approx\frac{C_\dyn}{N}.
926: \eeq
927: The exact dynamical entropy~(\ref{sigf})
928: and the refined a priori prediction~(\ref{sigref}) respectively yield
929: \beq
930: C_\dyn=4p\e^{-4p}(1-p+4p^2),\quad
931: C_\ap=4p\e^{-2p}(1-p\e^{-2p})(1-2p\e^{-2p}),
932: \eeq
933: i.e., for $p=1/2$,
934: $C_\dyn\approx0.406006$, $C_\ap\approx0.379540$.
935: 
936: \subsection{Correlations: Dynamics vs.~a priori ensemble}
937: 
938: Let us define the spin correlation function $C_n$
939: and the connected energy correlation function $G_n$ as
940: \beq
941: C_n=\dmean{\s_0\s_n},\quad
942: G_n=\dmean{\t_0\t_n}-E^2=\dmean{\s_0\s_1\s_n\s_{n+1}}-E^2.
943: \eeq
944: 
945: This section is devoted to a comparison between the exact
946: energy correlation function $G_n$
947: and the prediction of the refined a priori description.
948: It turns out that the spin correlation function $C_n$
949: is more difficult to handle,
950: both from the a priori and from the dynamical viewpoint.
951: 
952: The energy correlation function $(G_n)_\ap$
953: in the a priori ensemble at fixed energy density $E$
954: can again be evaluated by the transfer-matrix method.
955: We have, for~$n\ge0$ in the bulk of an infinitely long chain,
956: \beq
957: (G_n)_\ap
958: =\langle L_+\vert\En\vert R_-\rangle\langle L_-\vert\En\vert R_+\rangle
959: \left(\frac{\lam_-}{\lam_+}\right)^n.
960: \eeq
961: In this expression, $\En=\diag(-1,+1)$ is the energy operator, while
962: \beq
963: \langle L_\pm\vert=\frac{1}{\lam_\pm^2+1}\pmatrix{\lam_\pm&\e^\g},\quad
964: \vert R_\pm\rangle=\pmatrix{\e^\g\lam_\pm\cr1}
965: \eeq
966: are the left and right eigenvectors of $\T$
967: associated with the eigenvalues $\lam_\pm$.
968: We have consistently $\langle L_+\vert\En\vert R_+\rangle=E$.
969: After some algebra we obtain the following expression,
970: involving only the mean energy~$E$~\cite{ldt}:
971: \beq
972: (G_n)_\ap=(1-E^2)\left(-\frac{1+E}{1-E}\right)^n.
973: \label{cr}
974: \eeq
975: The connected energy correlation function in the a priori ensemble
976: thus exhibits an exponential fall-off, modulated by an oscillating sign.
977: 
978: The energy (i.e., occupancy, in the RSA language) correlation function
979: of the blocked configurations,
980: as they are generated by the zero-temperature dynamics,
981: has also been evaluated exactly~\cite{usrsa}.
982: Skipping again every detail, one has
983: \beq
984: (G_n)_\dyn
985: =2p\e^{-2p}\left((1-2p)\frac{(-2p)^n}{n!}
986: -2p\sum_{m\ge n+1}\frac{(-2p)^m}{m!}\right).
987: \label{cfinf}
988: \eeq
989: Whenever $p\ne1/2$, i.e., $T_0\ne\infty$,
990: the first term is leading, hence $(G_n)_\dyn\sim(-2p)^n/n!$.
991: In the case of an infinite initial temperature ($p=1/2$),
992: one has $(G_n)_\dyn\sim(-1)^n/(n+1)!$.
993: The connected energy correlations therefore exhibit a factorial decay,
994: modulated by an oscillating sign.
995: 
996: \section{Zero-temperature single-spin-flip dynamics in two dimensions:
997: Honeycomb vs.~square lattice}
998: \label{two}
999: 
1000: Leaving aside the one-dimensional realm,
1001: where exact analytical results are available,
1002: by means of an exact mapping of the dynamics onto an RSA problem,
1003: let us now turn to some novel results pertaining to two-dimensional examples
1004: of the zero-temperature single-spin-flip dynamics
1005: described in Section~\ref{gene}.
1006: We will investigate the following cases:
1007: 
1008: \begin{itemize}
1009: 
1010: \item
1011: {\it Honeycomb lattice}.
1012: The coordination number $z=3$ is odd,
1013: so that the standard zero-temperature dynamics~(\ref{wz}) is a descent dynamics.
1014: 
1015: \item 
1016: {\it Square lattice.}
1017: The coordination number $z=4$ is even,
1018: so that the situation is qualitatively similar to that of the chain.
1019: The dynamics~(\ref{wz}) is a bona fide coarsening dynamics
1020: for any non-zero value of the rate $W_0$,
1021: whereas the constrained dynamics ($W_0=0$) is a descent dynamics.
1022: 
1023: \end{itemize}
1024: 
1025: In the following we discuss in parallel
1026: the standard zero-temperature dynamics on the honeycomb lattice
1027: and the constrained one on the square lattice.
1028: Both dynamics lead to metastability:
1029: the system gets trapped in a finite time into one
1030: of many blocked configurations (zero-temperature metastable states).
1031: Figure~\ref{fighc} shows typical blocked configuration
1032: generated by both dynamics described above on samples of $150^2$ spins.
1033: 
1034: \begin{figure}[htb]
1035: \begin{center}
1036: \includegraphics[angle=90,width=.48\linewidth]{figh.eps}
1037: {\hskip 1mm}
1038: \includegraphics[angle=90,width=.48\linewidth]{figc.eps}
1039: \caption{\small
1040: Typical blocked configuration generated by the zero-temperature
1041: single-spin-flip dynamics considered here.
1042: Black (resp.~white) sites represent up (resp.~down) spins.
1043: Left: dynamics on the honeycomb lattice ($z=3$ is odd).
1044: Right: constrained dynamics on the square lattice ($z=4$ is even, $W_0=0$).}
1045: \label{fighc}
1046: \end{center}
1047: \end{figure}
1048: 
1049: \subsection{Distribution of the number of flips}
1050: 
1051: The constrained single-spin-flip dynamics on the Ising chain
1052: described in Section~\ref{chain} is fully irreversible,
1053: in the sense that every spin flips at most once
1054: during the whole history of the system,
1055: before a blocked configuration is reached.
1056: 
1057: In the present situation,
1058: although the two dynamics considered are descent dynamics in a global sense
1059: (the total energy decreases at each spin flip),
1060: a given spin can flip more than once.
1061: It turns out that the maximum number of flips is $M_\ca=2$
1062: for the constrained dynamics on the square lattice,
1063: and $M_\he=5$ for the dynamics on the honeycomb lattice.
1064: Skipping details,
1065: let us mention that these bounds can be derived by considering the number
1066: of unsatisfied bonds around the four squares,
1067: or the three hexagons, which surround a given spin.
1068: Figure~\ref{figflip} shows an example of a spin whose history
1069: contains this maximum number of flips in each case.
1070: The mean number of flips of a given spin remains however well below unity,
1071: at least for uncorrelated initial configurations.
1072: For dynamics on the honeycomb lattice,
1073: the mean number of flips is $\dmean{M}_\he\approx0.4244$,
1074: while the fraction of spins which flip five times
1075: is of order $p_\he(5)\sim10^{-8}$.
1076: For constrained dynamics on the square lattice,
1077: the mean number of flips is $\dmean{M}_\ca\approx0.2577$,
1078: while the fraction of spins which flip twice is $p_\ca(2)\approx0.00203$.
1079: 
1080: \begin{figure}[htb]
1081: \begin{center}
1082: \includegraphics[angle=90,width=.5\linewidth]{bc.eps}
1083: \vskip 8mm
1084: \includegraphics[angle=90,width=.5\linewidth]{bh.eps}
1085: \caption{\small
1086: Examples of histories such that a given spin $\s_0$
1087: experiences the maximum allowed number of flips.
1088: At each step the spin about to flip is circled.
1089: The letter F marks those steps where $\s_0$ is about to flip.
1090: Upper panel (left to right): square lattice:
1091: $\s_0$ (at center) flips $M_\ca=2$ times.
1092: Lower panel (top to bottom and left to right): honeycomb lattice:
1093: $\s_0$ (at bottom of uppermost hexagon) flips $M_\he=5$ times.}
1094: \label{figflip}
1095: \end{center}
1096: \end{figure}
1097: 
1098: \subsection{Distribution of the blocking time}
1099: 
1100: We have measured by means of numerical simulations
1101: the blocking time of finite samples of size $N\times N$
1102: for the descent dynamics on the honeycomb and square lattice described above,
1103: with uncorrelated random initial configurations.
1104: Figure~\ref{figtemps}
1105: shows a plot of the mean blocking time $\dmean{T_N}$ against the size
1106: (number of spins $N^2$) of the simulated samples.
1107: The data are well fitted by the form $\dmean{T_N}=a\ln N^2+b+c/N$.
1108: 
1109: This observation suggests that the late stages of the descent dynamics
1110: can again be described in an effective way
1111: by a dilute population of two-level systems,
1112: which relax independently of each other,
1113: with a common characteristic time $\t$,
1114: so that their density falls off exponentially, as $p(t)\sim\exp(-t/\t)$.
1115: Extreme-value statistics indeed implies
1116: \beq
1117: \dmean{T_N}\approx\t\ln N^2.
1118: \label{evs}
1119: \eeq
1120: The fits shown in Figure~\ref{figtemps} yield
1121: the characteristic times $\t_\he\approx2.49$ for the honeycomb lattice,
1122: and $\t_\ca\approx1.84$ for the square lattice.
1123: The fact that these estimates do not identify with simple numbers,
1124: together with the rather large observed correction to the result~(\ref{evs}),
1125: fitted by a $1/N$ term,
1126: suggests however that there might be more than one type of two-level systems.
1127: 
1128: \begin{figure}[htb]
1129: \begin{center}
1130: \includegraphics[angle=90,width=.5\linewidth]{figtemps.eps}
1131: \caption{\small
1132: Plot of the mean blocking time $\dmean{T_N}$
1133: of zero-temperature single-spin-flip dynamics,
1134: against the number of spins $N^2$.
1135: Full symbols: standard dynamics on the honeycomb lattice.
1136: Empty symbols: constrained dynamics on the square lattice.
1137: Full lines: fits (see text) with asymptotic slopes
1138: $\t_\he\approx2.49$ and $\t_\ca\approx1.84$.}
1139: \label{figtemps}
1140: \end{center}
1141: \end{figure}
1142: 
1143: \subsection{Spin correlations}
1144: 
1145: We have measured the on-axis spin correlation function
1146: $C_n=\dmean{\s_\vecz\s_\vecn}$,
1147: where $\vecn$ denotes the site at distance $n$ lattice bonds
1148: from the origin $\vecz$ along one of the main axes
1149: of either lattice, and where $\dmean{\dots}$
1150: again denotes an average over the blocked configurations.
1151: The spin correlation at unit distance $C_1$
1152: is nothing but minus the mean energy density
1153: $E_\infty$ of the blocked configurations per bond.
1154: We obtain $(C_1)_\he=-(E_\infty)_\he\approx0.7485$ for the honeycomb lattice,
1155: and $(C_1)_\ca=-(E_\infty)_\ca\approx0.6126$ for the square lattice.
1156: 
1157: \begin{figure}[htb]
1158: \begin{center}
1159: \includegraphics[angle=90,width=.5\linewidth]{figcor.eps}
1160: \caption{\small
1161: Logarithmic plot of the on-axis spin correlation function
1162: $C_n$ against distance $n$.
1163: Full symbols: standard dynamics on the honeycomb lattice.
1164: Empty symbols: constrained dynamics on the square lattice.
1165: Full straight lines: least-square fits with inverse slopes
1166: $\xi_\he\approx1.27$ and $\xi_\ca\approx0.79$.}
1167: \label{figcor}
1168: \end{center}
1169: \end{figure}
1170: 
1171: Figure~\ref{figcor}
1172: shows a logarithmic plot of the on-axis spin correlation function,
1173: against distance $n$.
1174: The data at large enough distances are well fitted by straight lines,
1175: demonstrating that correlations fall off exponentially to zero,
1176: as $C_n\sim\exp(-n/\xi)$.
1177: The inverse slopes of the least-square fits yield
1178: $\xi_\he\approx1.27$ for the honeycomb lattice,
1179: and $\xi_\ca\approx0.79$ for the square lattice.
1180: 
1181: \section{Discussion}
1182: \label{dis}
1183: 
1184: The main focus of the present paper is on zero-temperature
1185: single-spin-flip dynamics of ferromagnetic Ising models.
1186: We have emphasized that there are two kinds of such dynamics.
1187: The first kind ($\I$--type dynamics, in the classification of~\cite{ns})
1188: generically corresponds to bona fide coarsening dynamics,
1189: and lead to phase ordering by domain growth,
1190: whereas the second kind ($\F$--type dynamics)
1191: corresponds to descent dynamics,
1192: and provides interesting examples of dynamical systems with many attractors.
1193: 
1194: The one-dimensional case has been reviewed in Section~\ref{chain}.
1195: The existence of an exact mapping of the constrained zero-temperature dynamics
1196: onto the dimer RSA problem opens up the possibility
1197: of an exact analytical evaluation of many dynamical observables,
1198: including characteristics of the attractors~\cite{usrsa}.
1199: The comparison between the exact results thus obtained
1200: and the predictions of the a priori ensemble
1201: allows to test the validity of the so-called Edwards hypothesis,
1202: in a regime which is very distant from the situation of gentle tapping,
1203: considered originally by Edwards.
1204: On the one hand, a flat average over the a priori ensemble of attractors
1205: (especially in its refined version,
1206: defined by imposing the value of the observed mean energy density)
1207: provides inexact, but numerically reasonable,
1208: predictions for physical quantities.
1209: The numerical accuracy of a priori predictions may even by very good,
1210: like e.g.~for the dynamical entropy (see Figure~\ref{figent}).
1211: On the other hand, the energy correlation function provides an example
1212: of a qualitative discrepancy between the actual behavior of attractors
1213: and the prediction of any a priori ensemble.
1214: Connected correlations indeed exhibit a factorial fall-off in $1/n!$
1215: with distance, which is a generic characteristic of lattice RSA problems,
1216: whereas any flat measure over an a priori ensemble
1217: can be described by the transfer-matrix formalism,
1218: and therefore leads to exponentially decaying correlations in one dimension.
1219: Another example of a qualitative discrepancy concerns the lengths
1220: of ordered clusters in one-dimensional spin models.
1221: Cluster lengths are automatically statistically independent
1222: and exponentially distributed in a priori ensembles,
1223: again as a consequence of the underlying transfer-matrix formalism,
1224: whereas both of these properties have been shown to be violated
1225: in several examples of zero-temperature dynamics~\cite{bfs,uskawa}.
1226: 
1227: Examples of zero-temperature dynamics
1228: on two-dimensional Ising models have then been investigated numerically.
1229: The full dynamics on the honeycomb lattice
1230: and the constrained one on the square lattice exhibit very similar features.
1231: The observed logarithmic growth of the mean blocking time suggests that
1232: the late stages of both examples of descent dynamics
1233: are dominated by the relaxation of independent
1234: two-level systems, in analogy with the one-dimensional situation.
1235: This mechanism, in turn, is rather suggestive that attractors are likely
1236: to have a non-trivial statistics.
1237: The quantitative comparison with uniform measures on a priori ensembles
1238: is far more difficult than in one dimension,
1239: because a priori ensembles themselves may possess phase transitions
1240: as a function of their density.
1241: In particular, the accurately observed exponential fall-off
1242: of spin correlations is not very informative in that respect.
1243: Finally, it is worth noticing that the metastable states
1244: put forward in a recent work on competitive cluster growth~\cite{lume}
1245: share many of the features of the present two-dimensional attractors.
1246: 
1247: To sum up this investigation of the statistics of attractors
1248: of zero-temperature descent dynamics in finite dimensions,
1249: it turns out that
1250: the so-called Edwards hypothesis does not hold in this context in general.
1251: The predictions of the a priori approach, however,
1252: often provide good numerical approximations.
1253: This picture is corroborated by a recent work by Camia~\cite{camia}.
1254: 
1255: \Bibliography{99}
1256: 
1257: \bibitem{ange}
1258: Angell C A 1995 Science {\bf 267} 1924
1259: 
1260: \bibitem{glassy}
1261: Barrat J L, Dalibard D, Feigelman M, and Kurchan J eds 2003
1262: {\it Slow Relaxations and Non\-equi\-li\-brium Dynamics in Condensed Matter}
1263: Proceedings of Les Houches Summer School Session LXXVII (Berlin: Springer)
1264: 
1265: \bibitem{gold}
1266: Goldstein M 1969 J. Chem. Phys. {\bf 51} 3728
1267: 
1268: \bibitem{tap}
1269: Thouless D J, Anderson P W, and Palmer R G 1977 Phil. Mag. {\bf 35} 593
1270: 
1271: \bibitem{ks}
1272: Kirkpatrick K and Sherrington D 1978 Phys. Rev. B {\bf 17} 4384
1273: 
1274: \bibitem{gs}
1275: Gaveau B and Schulman L S 1998 J. Math. Phys. {\bf 39} 1517
1276: 
1277: \bibitem{bir}
1278: Biroli G and Monasson R 2000 Europhys. Lett. {\bf 50} 155
1279: \nonum
1280: Biroli G and Kurchan J 2001 Phys. Rev. E {\bf 64} 016101
1281: 
1282: \bibitem{ldt}
1283: Dean D S and Lef\`evre A 2001 Phys. Rev. Lett. {\bf 86} 5639
1284: \nonum
1285: Lef\`evre A and Dean D S 2001 J. Phys. A {\bf 34} L213
1286: \nonum
1287: Lef\`evre A 2002 J. Phys. A {\bf 35} 9037
1288: \nonum
1289: Dean D S and Lef\`evre A 2001 Phys. Rev. E {\bf 64} 046110
1290: 
1291: \bibitem{prbr}
1292: Prados A and Brey J J 2001 J. Phys. A {\bf 34} L453
1293: 
1294: \bibitem{usrsa}
1295: De Smedt G, Godr\`eche C, and Luck J M 2002 Eur. Phys. J. B {\bf 27} 363
1296: 
1297: \bibitem{gibbs}
1298: Gibbs J W 1906 {\it The Scientific Papers of J Willard Gibbs}
1299: (New York: Longmans, Green \& Co)
1300: \nonum
1301: Bumstead H A and Van Name R G eds 1961
1302: {\it The Scientific Papers of J Willard Gibbs} in 2 vols (New York: Dover)
1303: 
1304: \bibitem{ll}
1305: Landau L D and Lifshitz E M 1959 {\it Statistical Physics} (Oxford: Pergamon)
1306: \nonum
1307: Landau L D and Lifshitz E M 1960 {\it Electrodynamics of Continuous Media}
1308: (Oxford: Pergamon)
1309: 
1310: \bibitem{fisher}
1311: Fisher M E 1967 Physics {\bf 3} 255
1312: 
1313: \bibitem{langer}
1314: Langer J S 1967 Ann. Phys. {\bf 41} 108
1315: 
1316: \bibitem{frenkel}
1317: Frenkel J 1955 {\it Kinetic Theory of Liquids} (New York: Dover)
1318: 
1319: \bibitem{zinn}
1320: Zinn-Justin J 1989 {\it Quantum Field Theory and Critical Phenomena}
1321: (Oxford: Clarendon)
1322: 
1323: \bibitem{bustillos}
1324: Ticona Bustillos A, Heermann D W, and Cordeiro C E 2004 J. Chem. Phys.
1325: {\bf 121} 4804
1326: 
1327: \bibitem{sconf}
1328: J\"ackle J 1981 Phil. Mag. B {\bf 44} 533
1329: \nonum
1330: Palmer R G 1982 Adv. Phys. {\bf 31} 669
1331: 
1332: \bibitem{ktw}
1333: Kirkpatrick T R and Wolynes P G 1987 Phys. Rev. A {\bf 35} 3072
1334: \nonum
1335: Kirkpatrick T R and Thirumalai D 1987 Phys. Rev. B {\bf 36} 5388
1336: \nonum
1337: Kirkpatrick T R and Wolynes P G 1987 Phys. Rev. B {\bf 36} 8552
1338: \nonum
1339: Thirumalai D and Kirkpatrick T R 1988 Phys. Rev. B {\bf 38} 4881
1340: 
1341: \bibitem{sw}
1342: Stillinger F H and Weber T A 1982 Phys. Rev. A {\bf 25} 978
1343: \nonum
1344: Stillinger F H and Weber T A 1984 Science {\bf 225} 983
1345: 
1346: \bibitem{fv}
1347: Franz S and Virasoro M A 2000 J. Phys. A {\bf 33} 891
1348: 
1349: \bibitem{bm}
1350: Berg J and Mehta A 2001 Europhys. Lett. {\bf 56} 784
1351: \nonum
1352: Berg J and Mehta A 2001 Adv. Complex Syst. {\bf 4} 309
1353: 
1354: \bibitem{pbt}
1355: Prados A and Brey J J 2001 Phys. Rev. E {\bf 66} 041308
1356: 
1357: \bibitem{edwards}
1358: Edwards S F 1994 in {\it Granular Matter: An Interdisciplinary Approach}
1359: Mehta A ed (New York: Springer)
1360: 
1361: \bibitem{fdt}
1362: Cugliandolo L F and Kurchan J 1994 J. Phys. A {\bf 27} 5749
1363: \nonum
1364: Cugliandolo L F, Kurchan J, and Parisi G 1994 J. Phys. I (France) {\bf 4} 1641
1365: \nonum
1366: Cugliandolo L F, Kurchan J, and Peliti L 1997 Phys. Rev. E {\bf 55} 3898
1367: 
1368: \bibitem{ba}
1369: Barrat A, Kurchan J, Loreto V, and Sellitto M 2001 Phys. Rev. Lett. {\bf 85}
1370: 5034
1371: \nonum
1372: Barrat A, Kurchan J, Loreto V, and Sellitto M 2001 Phys. Rev. E {\bf 63} 051301
1373: 
1374: \bibitem{red}
1375: Spirin V, Krapivsky P L, and Redner S 2001 Phys. Rev. E {\bf 63} 036118
1376: \nonum
1377: Spirin V, Krapivsky P L, and Redner S 2002 Phys. Rev. E {\bf 65} 016119
1378: 
1379: \bibitem{ns}
1380: Nanda S, Newman C M, and Stein D L 2000 in {\it On Dobrushin's Way
1381: (from Probability Theory to Statistical Physics)} Dubrushin R L \etal eds
1382: Amer. Math. Soc. Transl. {\bf 198}
1383: \nonum
1384: Newman C M and Stein D L 2000 Physica A {\bf 279} 159
1385: \nonum
1386: Nanda S, Newman C M, and Stein D L 2004 preprint cond-mat/0411286
1387: 
1388: \bibitem{barc}
1389: Sollich P and Ritort F eds 2002 {\it Workshop on glassy behaviour
1390: of kinetically constrained models} J.~Phys. Cond. Matt. {\bf 14} 1381-1696
1391: 
1392: \bibitem{riso}
1393: Ritort F and Sollich P 2003 Adv. Phys. {\bf 52} 219
1394: 
1395: \bibitem{glau}
1396: Glauber R J 1963 J. Math. Phys. {\bf 4} 294
1397: 
1398: \bibitem{bray}
1399: Bray A J 1994 Adv. Phys. {\bf 43} 357
1400: 
1401: \bibitem{flory}
1402: Flory J P 1939 J. Am. Chem. Soc. {\bf 61} 1518
1403: 
1404: \bibitem{core}
1405: Cohen E R and Reiss H 1963 J. Chem. Phys. {\bf 38} 680
1406: 
1407: \bibitem{rsa}
1408: Evans J W 1993 Rev. Mod. Phys. {\bf 65} 1281
1409: 
1410: \bibitem{gumbel}
1411: Gumbel E J 1958 {\it Statistics of Extremes} (New York:
1412: Columbia University Press).
1413: 
1414: \bibitem{cri}
1415: Crisanti A, Ritort F, Rocco A, and Sellitto M 2000 J. Chem. Phys. {\bf 113}
1416: 10615
1417: 
1418: \bibitem{bfs}
1419: Berg J, Franz S, and Sellitto M 2002 Eur. Phys. J. B {\bf 26} 349
1420: 
1421: \bibitem{baxter}
1422: Baxter R J 1982 {\it Exactly Solved Models in Statistical Mechanics}
1423: (London: Academic)
1424: 
1425: \bibitem{tm}
1426: Crisanti A, Paladin G, and Vulpiani A 1992
1427: {\it Products of Random Matrices in Statistical Physics}
1428: (Springer, Berlin, 1992)
1429: \nonum
1430: Luck J M 1992 {\it Syst\`emes d\'esordonn\'es unidimensionnels}
1431: in French (Saclay: Collection Al\'ea)
1432: 
1433: \bibitem{uskawa}
1434: De Smedt G, Godr\`eche C, and Luck J M 2003 Eur. Phys. J. B {\bf 32} 215
1435: 
1436: \bibitem{lume}
1437: Luck J M and Mehta A 2004 preprint cond-mat/0410385
1438: 
1439: \bibitem{camia}
1440: Camia F 2004 preprint cond-mat/0410543
1441: 
1442: \endbib
1443: \end{document}
1444: