1: \documentclass[pre,aps,superscriptaddress,floatfix,twocolumn,showpacs]{revtex4}
2:
3: \usepackage{graphicx}
4: \usepackage{amsmath}
5: \voffset=1cm
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: %%%%%%%%%%%%%%%%%%%%%%%%% DEFINIZIONI DI FONT %%%%%%%%%%%%%%%%%%%%%%%%%%%
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: \font\tenmib=cmmib10
10: \font\sevenmib=cmmib10 scaled 800
11: \font\titolo=cmbx12
12: \font\titolone=cmbx10 scaled\magstep 2
13: \font \titolino=cmbx10
14: \font\cs=cmcsc10
15: \font\sc=cmcsc10
16: \font\css=cmcsc8
17: \font\ss=cmss10
18: \font\sss=cmss8
19: \font\crs=cmbx8
20: \font\ninerm=cmr9
21: \font\ottorm=cmr8
22: \textfont5=\tenmib
23: \scriptfont5=\sevenmib
24: \font\euftw=eufm9 scaled\magstep1
25: \font\euftww=eufm7 scaled\magstep1
26: \font\euftwww=eufm5 scaled\magstep1
27: \font\msytw=msbm9 scaled\magstep1
28: \font\msytww=msbm7 scaled\magstep1
29: \font\msytwww=msbm5 scaled\magstep1
30: \font\msytwwww=msbm4 scaled\magstep1
31: \font\indbf=cmbx10 scaled\magstep2
32: \font\type=cmtt10
33: \def\st{\scriptstyle}
34:
35: \font\ottorm=cmr8\font\ottoi=cmmi8\font\ottosy=cmsy8%
36: \font\ottobf=cmbx8\font\ottott=cmtt8%
37: \font\ottocss=cmcsc8%
38: \font\ottosl=cmsl8\font\ottoit=cmti8%
39: \font\sixrm=cmr6\font\sixbf=cmbx6\font\sixi=cmmi6\font\sixsy=cmsy6%
40: \font\fiverm=cmr5\font\fivesy=cmsy5
41: \font\fivei=cmmi5
42: \font\fivebf=cmbx5%
43:
44: \def\ottopunti{\def\rm{\fam0\ottorm}%
45: \textfont0=\ottorm\scriptfont0=\sixrm\scriptscriptfont0=\fiverm%
46: \textfont1=\ottoi\scriptfont1=\sixi\scriptscriptfont1=\fivei%
47: \textfont2=\ottosy\scriptfont2=\sixsy\scriptscriptfont2=\fivesy%
48: \textfont3=\tenex\scriptfont3=\tenex\scriptscriptfont3=\tenex%
49: \textfont4=\ottocss\scriptfont4=\sc\scriptscriptfont4=\sc%
50: %\scriptfont4=\ottocss\scriptscriptfont4=\ottocss%
51: \textfont5=\tenmib\scriptfont5=\sevenmib\scriptscriptfont5=\fivei
52: \textfont\itfam=\ottoit\def\it{\fam\itfam\ottoit}%
53: \textfont\slfam=\ottosl\def\sl{\fam\slfam\ottosl}%
54: \textfont\ttfam=\ottott\def\tt{\fam\ttfam\ottott}%
55: \textfont\bffam=\ottobf\scriptfont\bffam=\sixbf%
56: \scriptscriptfont\bffam=\fivebf\def\bf{\fam\bffam\ottobf}%
57: %\tt\ttglue=.5em plus.25em minus.15em%
58: \setbox\strutbox=\hbox{\vrule height7pt depth2pt width0pt}%
59: \normalbaselineskip=9pt\let\sc=\sixrm\normalbaselines\rm}
60: \let\nota=\ottopunti%
61:
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: %%%%%%%%%%%%%%%%%%%%%% SIMBOLI VARI %%%%%%%%%%%%%%%%%%%%%%%
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65:
66: \let\a=\alpha \let\b=\beta \let\g=\gamma \let\d=\delta \let\e=\varepsilon
67: \let\z=\zeta \let\h=\eta \let\th=\theta \let\k=\kappa \let\l=\lambda
68: \let\m=\mu \let\n=\nu \let\x=\xi \let\p=\pi \let\r=\rho
69: \let\s=\sigma \let\t=\tau \let\f=\varphi \let\ph=\varphi\let\c=\chi
70: \let\ps=\psi \let\y=\upsilon \let\o=\omega\let\si=\varsigma
71: \let\G=\Gamma \let\D=\Delta \let\Th=\Theta\let\L=\Lambda \let\X=\Xi
72: \let\P=\Pi \let\Si=\Sigma \let\F=\Phi \let\Ps=\Psi
73: \let\O=\Omega \let\Y=\Upsilon
74:
75: \def\\{\hfill\break} \let\==\equiv
76: \let\txt=\textstyle\let\dis=\displaystyle
77:
78: \let\io=\infty \def\Dpr{\V\dpr\,}
79: \def\aps{{\it a posteriori\ }}\def\ap{{\it a priori\ }}
80: \let\0=\noindent\def\pagina{{\vfill\eject}}
81: \def\bra#1{{\langle#1|}}\def\ket#1{{|#1\rangle}}
82: \def\media#1{{\langle#1\rangle}}
83: %\def\ie{\hbox{\it i.e.\ }}\def\eg{\hbox{\it e.g.\ }}
84: \def\ie{{i.e. }}\def\eg{{e.g. }}
85: \let\dpr=\partial \def\der{{\rm d}} \let\circa=\cong
86: \def\arccot{{\rm arccot}}
87:
88: \def\tende#1{\,\vtop{\ialign{##\crcr\rightarrowfill\crcr
89: \noalign{\kern-1pt\nointerlineskip} \hskip3.pt${\scriptstyle
90: #1}$\hskip3.pt\crcr}}\,}
91: \def\circage{\lower2pt\hbox{$\,\buildrel > \over {\scriptstyle \sim}\,$}}
92: \def\otto{\,{\kern-1.truept\leftarrow\kern-5.truept\to\kern-1.truept}\,}
93: \def\fra#1#2{{#1\over#2}}
94:
95: \def\PPP{{\cal P}}\def\EE{{\cal E}}\def\MM{{\cal M}} \def\VV{{\cal V}}
96: \def\CC{{\cal C}}\def\FF{{\cal F}} \def\HHH{{\cal H}}\def\WW{{\cal W}}
97: \def\TT{{\cal T}}\def\NN{{\cal N}} \def\BBB{{\cal B}}\def\II{{\cal I}}
98: \def\RR{{\cal R}}\def\LL{{\cal L}} \def\JJ{{\cal J}} \def\OO{{\cal O}}
99: \def\DD{{\cal D}}\def\AAA{{\cal A}}\def\GG{{\cal G}} \def\SS{{\cal S}}
100: \def\KK{{\cal K}}\def\UU{{\cal U}} \def\QQ{{\cal Q}} \def\XXX{{\cal X}}
101:
102: \def\T#1{{#1_{\kern-3pt\lower7pt\hbox{$\widetilde{}$}}\kern3pt}}
103: \def\VVV#1{{\underline #1}_{\kern-3pt
104: \lower7pt\hbox{$\widetilde{}$}}\kern3pt\,}
105: \def\W#1{#1_{\kern-3pt\lower7.5pt\hbox{$\widetilde{}$}}\kern2pt\,}
106: \def\Re{{\rm Re}\,}\def\Im{{\rm Im}\,}
107: \def\lis{\overline}\def\tto{\Rightarrow}
108: \def\etc{{\it etc}} \def\acapo{\hfill\break}
109: \def\mod{{\rm mod}\,} \def\per{{\rm per}\,} \def\sign{{\rm sign}\,}
110: \def\indica{\leaders \hbox to 0.5cm{\hss.\hss}\hfill}
111: \def\guida{\leaders\hbox to 1em{\hss.\hss}\hfill}
112:
113: \def\hh{{\bf h}} \def\HH{{\bf H}} \def\AA{{\bf A}} \def\qq{{\bf q}}
114: \def\BB{{\bf B}} \def\XX{{\bf X}} \def\PP{{\bf P}} \def\pp{{\bf p}}
115: \def\vv{{\bf v}} \def\xx{{\bf x}} \def\yy{{\bf y}} \def\zz{{\bf z}}
116: \def\aaa{{\bf a}}\def\bbb{{\bf b}}\def\hhh{{\bf h}}\def\III{{\bf I}}
117:
118: \def\ul{\underline}
119: \def\olu{{\overline{u}}}
120:
121: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
122: %%%%%%%%%%%%%%%%% LETTERE GRECHE E LATINE IN NERETTO %%%%%%%%%%%%%%%%%
123: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124:
125: % lettere greche e latine in neretto italico - pag.430 del manuale
126: \mathchardef\aa = "050B
127: \mathchardef\bb = "050C
128: \mathchardef\ggg = "050D
129: \mathchardef\xxx = "0518
130: %\mathchardef\hhh = "0511
131: \mathchardef\zzzzz= "0510
132: \mathchardef\oo = "0521
133: \mathchardef\lll = "0515
134: \mathchardef\mm = "0516
135: \mathchardef\Dp = "0540
136: \mathchardef\H = "0548
137: \mathchardef\FFF = "0546
138: \mathchardef\ppp = "0570
139: \mathchardef\nn = "0517
140: \mathchardef\ff = "0527
141: \mathchardef\pps = "0520
142: %\mathchardef\XXX = "0504
143: \mathchardef\FFF = "0508
144: \mathchardef\nnnnn= "056E
145:
146: \def\ol#1{{\overline #1}}
147:
148: \def\to{\rightarrow}
149: \def\la{\left\langle}
150: \def\ra{\right\rangle}
151:
152: \def\Overline#1{{\bar#1}}
153: \let\ciao=\bye
154: \def\qed{\raise1pt\hbox{\vrule height5pt width5pt depth0pt}}
155: \def\hf#1{{\hat \f_{#1}}}
156: \def\barf#1{{\tilde \f_{#1}}} \def\tg#1{{\tilde g_{#1}}}
157: \def\bq{{\bar q}}
158: %\def\Val{{\bf Val}}
159: \def\Val{{\rm Val}}
160: \def\indic{\hbox{\raise-2pt \hbox{\indbf 1}}}
161:
162: \def\RRR{\hbox{\msytw R}} \def\rrrr{\hbox{\msytww R}}
163: \def\rrr{\hbox{\msytwww R}} \def\CCC{\hbox{\msytw C}}
164: \def\cccc{\hbox{\msytww C}} \def\ccc{\hbox{\msytwww C}}
165: \def\NNN{\hbox{\msytw N}} \def\nnnn{\hbox{\msytww N}}
166: \def\nnn{\hbox{\msytwww N}} \def\ZZZ{\hbox{\msytw Z}}
167: \def\zzzz{\hbox{\msytww Z}} \def\zzz{\hbox{\msytwww Z}}
168: \def\TTT{\hbox{\msytw T}} \def\tttt{\hbox{\msytww T}}
169: \def\ttt{\hbox{\msytwww T}}
170: \def\QQQ{\hbox{\msytw Q}} \def\qqqq{\hbox{\msytww Q}}
171: \def\qqq{\hbox{\msytwww Q}}
172:
173: \def\vvv{\hbox{\euftw v}} \def\vvvv{\hbox{\euftww v}}
174: \def\vvvvv{\hbox{\euftwww v}}\def\www{\hbox{\euftw w}}
175: \def\wwww{\hbox{\euftww w}} \def\wwwww{\hbox{\euftwww w}}
176: \def\vvr{\hbox{\euftw r}} \def\vvvr{\hbox{\euftww r}}
177:
178: \def\ul#1{{\underline#1}}
179: \def\Sqrt#1{{\sqrt{#1}}}
180: %\def\Sqrt#1{{{#1}^{\fra{1}{2}}}}
181: \def\V0{{\bf 0}}
182: \def\defi{\,{\buildrel def\over=}\,}
183: \def\lhs{{\it l.h.s.}\ }
184: \def\rhs{{\it r.h.s.}\ }
185: \def\defin{{\buildrel def\over=}}
186: \def\V#1{{\underline#1}}
187: \def\Prob{{\rm Prob}}
188: \newcommand{\beq}{\begin{equation}}
189: \newcommand{\eeq}{\end{equation}}
190: \newcommand{\wh}{\widehat}
191: \newcommand{\wt}{\widetilde}
192:
193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
195:
196: \begin{document}
197:
198: \title{
199: Fluctuation Relation
200: beyond Linear Response Theory
201: }
202: \author{
203: A.~Giuliani\footnote{e-mail: alessandro.giuliani@roma1.infn.it}
204: }
205: \affiliation{
206: Dipartimento di Fisica, Universit\`a di Roma
207: {\em La Sapienza}, P. A. Moro 2, 00185 Roma, Italy
208: }
209: \affiliation{
210: INFN, Universit\`a di Roma
211: {\em La Sapienza}, P. A. Moro 2, 00185 Roma, Italy
212: }
213: \author{
214: F.~Zamponi\footnote{e-mail: francesco.zamponi@phys.uniroma1.it}
215: }
216: \affiliation{
217: Dipartimento di Fisica, Universit\`a di Roma
218: {\em La Sapienza}, P. A. Moro 2, 00185 Roma, Italy
219: }
220: \affiliation{
221: INFM -- CRS Soft, Universit\`a di Roma
222: {\em La Sapienza}, P. A. Moro 2, 00185 Roma, Italy
223: }
224: \author{
225: G.~Gallavotti\footnote{e-mail: giovanni.gallavotti@roma1.infn.it}
226: }
227: \affiliation{
228: Dipartimento di Fisica, Universit\`a di Roma
229: {\em La Sapienza}, P. A. Moro 2, 00185 Roma, Italy
230: }
231: \affiliation{
232: INFN, Universit\`a di Roma
233: {\em La Sapienza}, P. A. Moro 2, 00185 Roma, Italy
234: }
235: \date{\today}
236: \begin{abstract}
237: The Fluctuation Relation (FR) is an asymptotic result on the distribution of
238: certain observables averaged over time intervals $\t$ as $\t\to\io$
239: and it is a generalization of the fluctuation--dissipation
240: theorem to far from equilibrium systems in a steady state which
241: reduces to the usual Green--Kubo (GK) relation in the limit of small
242: external non conservative forces. FR is a theorem
243: for smooth uniformly hyperbolic systems, and it is assumed to be true
244: in all dissipative ``chaotic enough'' systems in a steady state.
245: In this paper we develop a theory of finite time
246: corrections to FR, needed to compare the asymptotic prediction
247: of FR with numerical observations, which necessarily involve
248: fluctuations
249: of observables averaged over finite time intervals $\t$.
250: We perform
251: a numerical test of FR in two cases in which non Gaussian fluctuations are
252: observable while GK does not apply and we get a non trivial verification of
253: FR that is
254: {\it independent of} and {\it different from} linear response theory.
255: Our results are compatible with the theory
256: of finite time corrections to FR, while FR would be {\it observably
257: violated}, well within the precision of our experiments, if such
258: corrections were neglected.
259: \end{abstract}
260: \pacs{05.40.-a,05.45.-a,05.45.Pq}
261: \maketitle
262:
263:
264: %%%%%%%%%%%%%%% TEXT %%%%%%%%%%%%%%%%
265:
266: Keywords: entropy production rate, fluctuation theorem,
267: non-Gaussian fluctuations, Green-Kubo relations
268:
269: \section{Introduction}
270:
271: \noindent
272: {\it Anosov systems and the fluctuation theorem} -
273: The fluctuation theorem concerns fluctuations of phase space
274: contraction in reversible hyperbolic (Anosov) systems. If time evolution is
275: described by a differential equation on phase space $M$: $\dot
276: x=X(x),\, x\in M$,
277: or by a map $S: x\rightarrow S(x)$ of $M$ one defines the {\it phase space
278: contraction} as, respectively, $\sigma(x)=-{\rm \,div\,} X(x)$ or
279: $\sigma(x)=-\log |\det\partial_x S(x)|$. Reversibility means that there is a
280: metric--preserving map $I$ of $M$ such that $IS=S^{-1}I$ if $S$ is the
281: time evolution over a certain time $t$ ({\it e.g.} $t=1$).
282: {\it If the system is Anosov}, that is if $M$ is compact and $S$ is smooth
283: and uniformly hyperbolic, see \cite{Ga99,GC95a,GC95b,Ga95c,Ru99},
284: the points $x$ will have a well defined
285: SRB distribution $\mu_{srb}$, \cite{Ru99}, {\it i.e.} almost all points
286: w.r.t. the volume measure will evolve
287: so that {\it all smooth observables} will have a well defined average
288: equal to the integral over the SRB distribution. Hence, in particular,
289: the time average of the function $\sigma(x)$
290: will be asymptotically given by the spatial average w.r.t.
291: the SRB distribution. In the case of discrete time maps:
292: \beq
293: \label{1}
294: \sigma_+\defin\lim_{\t\to\io} \frac1\t \sum_{j=0}^{\tau-1} \sigma(S^j(x))
295: =\int_M \sigma\;d\mu_{srb} \defin \langle\; \sigma \;\rangle_{srb}\eeq
296: If $\sigma_+ > 0$, let:
297: \beq
298: \label{2}
299: p(x)=\frac{1}{\tau \sigma_+} \sum_{j=0}^{\tau-1} \sigma(S^j(x))
300: \eeq
301: Analogous definitions are given
302: in the continuous time case.
303: The function $p(x)$ will have average $\langle\; p\;\rangle_{srb} = 1$
304: and distribution $\pi_\tau(dp)$ such that
305: \beq
306: \label{3}
307: \pi_\tau(\{p\in \Delta\})= e^{\tau \max_{p\in\Delta}\zeta_\infty(p)+ o(\t)}\;,
308: \eeq
309: where the correction at the exponent is $o(\t)$ w.r.t. $\t$ as $\t\to\io$.
310: The following {\it Fluctuation Relation}, discovered
311: in a numerical simulation
312: in \cite{ECM93} and formulated as a theorem for Anosov systems in
313: \cite{GC95a}, holds:
314: \beq
315: \label{4}
316: \zeta_\infty(p)=\zeta_\infty(-p)+p \sigma_+ \qquad{\rm for\ all}\ |p|<p^*
317: \eeq
318: where $\infty> p^*\ge1$ is a suitable (model dependent) constant that, in
319: general, is {\it different} from the maximum over $\tau$ and $x$ of
320: $p(x)$; note also that Eq.~\ref{4} is
321: (strictly speaking) meaningless in the {\it equilibrium
322: cases} in which the system is Hamiltonian and reversibility is the
323: usual velocity sign change: because of the division by $\sigma_+=0$ in
324: Eq.~\ref{2}.
325:
326: \vskip10pt
327:
328: \noindent
329: {\it The chaotic hypothesis} - Hyperbolicity is a paradigm
330: for disordered systems similar to the
331: small oscillations paradigm used for ordered motions: it does not
332: hold exactly in essentially all the physically interesting
333: systems. The {\it Chaotic Hypothesis} \cite{Ga99,GC95a,GC95b,Ga04,Ru04}
334: is that nevertheless
335: one can assume that chaotic motions
336: (in the sense of motions with at least one positive Lyapunov exponent)
337: exhibit some average properties of truly
338: hyperbolic motions. This hypothesis is a natural generalization of
339: the ergodic hypothesis,
340: \ie of the assumption that systems of many particles at equilibrium
341: are well
342: described on average by the microcanonical (or by the Gibbs)
343: distribution, even if they are not
344: really (or they are not proven to be) ergodic.
345: A consequence of the Chaotic Hypothesis is that
346: (dissipative) deterministic chaotic reversible motions
347: should have fluctuations of phase space contraction satisfying Eq.~\ref{4}.
348:
349: One interesting example of such motions is given by a system of
350: $N$ interacting particles
351: in $d$ dimensions subjected to nonconservative forces and kept in a
352: stationary state by a {\it reversible mechanical thermostat}.
353: It will be defined by a differential equation $\dot x=X_E(x)$ where
354: $x=(\dot{\ul q},\ul q) \in R^{2dN} \equiv M$ ({\it phase space})
355: and
356: \beq
357: \label{INTRO1}
358: m \ddot{\ul q} = \ul f(\ul q) + \ul g_E (\ul q) - \ul
359: \theta_E (\dot{\ul q},\ul q)
360: \eeq
361: where $m$ is the mass of the particles, $\ul f(\ul q)$ describes
362: the internal (conservative)
363: forces between the particles and $\ul g_E (\ul q)$ represents
364: the nonconservative
365: ``external'' force acting on the system. Finally,
366: $\ul \theta_E (\dot{\ul q},\ul q)$ is a
367: mechanical force that prevents the system from acquiring energy indefinitely:
368: this is why we shall
369: call it a {\it mechanical thermostat}.
370: Systems belonging to this class are frequently
371: used as microscopic models to describe
372: nonequilibrium stationary states induced by the application of a driving force
373: (temperature or velocity gradients, electric fields, etc.) on a
374: fluid system in contact with a thermal bath, \cite{EM90,Ru04}.
375: In this context, the phase space contraction
376: rate $\sigma(x)$ has been identified (setting $k_B=1$)
377: with the {\it entropy production rate} \cite{ECM93,Ga99,Ga04,Ru04},
378: the variable $p(x)$ is defined as
379: \beq
380: \label{pdix}
381: p(x) = \frac{1}{\tau \sigma_+} \int_0^\tau dt \ \sigma(S_t x)
382: \eeq
383: (where $x(t) \equiv S_t x$ is the solution of Eq.~\ref{INTRO1}
384: with initial datum $x(0)=x$)
385: and the fluctuation relation has been successfully tested in
386: several numerical simulations
387: \cite{ECM93,BGG97,BCL98,BPV98,GP99,GRS04,ZRA04}.
388: Having defined the notion of entropy production rate one can
389: define a ``duality'' between
390: fluxes $\ul J$ and forces $\ul E$ using $\sigma(x)$ as a
391: ``Lagrangian'' \cite{Ga04}:
392: \beq
393: \label{Jdef}
394: \ul J(\ul E,x) = \frac{\partial \sigma(x)}{\partial \ul E}
395: \eeq
396: In the limit $\ul E \rightarrow \ul 0$, {\it i.e.} close to equilibrium,
397: the fluctuation relation
398: leads to Onsager's reciprocity and to Green-Kubo's formulas for
399: transport coefficients
400: \cite{GR97,Ga96}:
401: \beq
402: \label{GK}
403: \mu_{ij} \equiv \lim_{\ul E \rightarrow \ul 0} \frac{\langle J_i
404: \rangle_{\ul E}}{E_j} =
405: \int_0^\infty dt \ \langle J_i(t) J_j(0) \rangle_{\ul E=\ul 0}
406: \eeq
407:
408: \vskip10pt
409: \noindent
410: {\it Gaussian distributions} -
411: If the distribution $\pi_\tau(p)$ is Gaussian,
412: $\pi_\tau(p) \propto \exp \left[ -\tau \frac{(p-1)^2}{2 \delta_\tau^2}\right]$,
413: from the fluctuation relation one can derive an {\it extension}
414: of the Green-Kubo relation, i.e. of
415: Eq.~\ref{GK}, to finite forces.
416:
417: Indeed, the fluctuation relation for a Gaussian distribution implies
418: that the dispersion $\d_\io^2$ of $p$ around its average (equal to $1$) is
419: $\delta_\infty^2 = 2/\s_+$ which is, in such case, an extension of a
420: Green--Kubo formula to non zero fields. One sees this by considering, for
421: instance, cases in which $\sigma(x)$ is linear in $E$ (as it will be in
422: the cases that we study numerically below).
423: Using time-translation invariance one can show that
424: \beq
425: \label{deltainfinito}
426: \delta^2_\infty = \frac{2}{\sigma_+^2}\int_0^\infty dt \
427: \langle (\sigma(t)-\sigma_+)(\sigma(0)-\sigma_+) \rangle_{E}
428: \eeq
429: and from the fluctuation relation $\delta_\infty^2 \sigma_+ = 2$
430: \beq
431: \sigma_+ = \int_0^\infty dt \
432: \langle (\sigma(t)-\sigma_+)(\sigma(0)-\sigma_+) \rangle_{E}
433: \eeq
434: Substituting $\sigma(t)=E J(t)$ in the latter expression, one
435: obtains the relation
436: \beq
437: \label{GKestesa}
438: \frac{\langle J \rangle_E}{E} = \int_0^\infty dt \
439: [\langle J(t) J(0) \rangle_E - \langle J \rangle_E^2]
440: \eeq
441: valid, {\it subject to the Gaussian assumption},
442: also for $E\ne0$.
443:
444: The leading order in $E$ of the latter relation
445: ({\it linear response}) is the Green-Kubo formula
446: for the equilibrium transport coefficient, Eq.~\ref{GK}.
447:
448: \vskip10pt
449: \noindent
450: {\it Numerical verification of the chaotic hypothesis} -
451: The simplest check of the applicability of the Chaotic Hypothesis
452: is a check of the fluctuation relation: of course even if the
453: check has a positive result this will not ``prove'' the hypothesis but
454: it will at least add confidence to it. A rather stringent test of the
455: fluctuation relation would be a check which
456: {\it cannot be reduced to a kind of Green-Kubo relation}; this requires
457: at least one of the two following conditions to be satisfied:
458: \begin{enumerate}
459: \item the distribution $\pi_\tau(p)$ is distinguishable from a Gaussian, or
460: \item deviations from the leading order in $E$ in Eq.~\ref{GKestesa},
461: {\it i.e.},
462: deviations from the Green-Kubo relation, are observed.
463: \end{enumerate}
464: This is very hard to obtain in numerical simulations of
465: Eq.~\ref{INTRO1} for the following
466: reasons:
467: \begin{enumerate}
468: \item to observe deviations from linearity in Eq.~\ref{GKestesa} one has to
469: apply very large forces $E$, then $\sigma_+$ is very large and
470: it becomes very difficult
471: to observe the negative values of $p(x)$ that are needed to
472: compute $\zeta_\infty(-p)$ in Eq.~\ref{4};
473: \item deviations from Gaussianity in $\pi_\tau(p)$
474: are observed only for values of $p$ that
475: differ significantly (of the order of $2 \delta_\io$) from $1$ and,
476: again, it is very difficult
477: to observe such values of $p$.
478: \end{enumerate}
479: Due to the limited computational resources available in the past decade,
480: all numerical computations that can be found in the literature
481: on systems described by Eq.~\ref{INTRO1} found that the measured
482: distribution $\pi_\tau(p)$ could not be
483: distinguished from a Gaussian distribution
484: in the interval of $p$ accessible to the numerical experiment
485: \cite{ECM93,BGG97,BCL98,ZRA04}.
486:
487: \vskip10pt
488:
489: The purpose of the present paper is to test the fluctuation relation,
490: in a numerical simulation
491: of a system described by Eq.~\ref{INTRO1}, for large
492: applied force when deviations from linearity can be observed, and
493: the distribution $\pi_\tau(p)$ is appreciably non-Gaussian.
494: This has become possible
495: thanks to the fast increase of computational power in the last decade. However,
496: it is still very difficult to reach values of $\tau$ which
497: can be confidently regarded as ``close'' to the asymptotic limit
498: $\tau \rightarrow \infty$;
499: thus to interpret our results we develop a theory of the $o(1)$
500: corrections to the function
501: $\zeta_\infty(p)$ in order to extract the limiting function
502: $\zeta_\infty(p)$ from the
503: numerical data. Taking into account the latter finite time corrections,
504: we successfully test the fluctuation relation for
505: non--Gaussian distributions and beyond the
506: linear response theory.
507:
508: The paper is organized as follows: section II is devoted to the theory
509: of finite $\tau$
510: corrections to the large deviation function
511: $\zeta_\infty(p)$ that is needed in the analysis of
512: the data; in section III we present the model and the details of the numerical
513: simulation; in section IV we present the details of the data analysis; finally,
514: in section V and VI we report the result of our simulations.
515:
516: \section{Finite time corrections to the Fluctuation Relation}
517: \label{sec:II}
518:
519: In the present section we
520: describe a strategy to study (in principle constructively)
521: the $O(1)$ corrections
522: in the exponent of Eq.~\ref{3}. The theory we propose will
523: hold {\it assuming that the time evolution is hyperbolic}
524: so that it can be applied to
525: physical systems only if the chaotic hypothesis is accepted. For
526: simplicity we consider only the case of discrete time evolution via a
527: map $S$.
528:
529: \subsection{SRB measure, symbolic dynamics and statistical mechanics}
530:
531: We study the distribution of $p$ at fixed $\tau$ via its Laplace
532: transform ({\it characteristic function}) $z_\tau(\lambda)$:
533: \beq
534: \label{FTC2.1}
535: z_\tau(\lambda) = -\frac{1}{\tau} \log
536: \langle e^{-\lambda \sum_{j=0}^{\tau-1}\sigma(S^jx)}\rangle_{srb}
537: \eeq
538: The main consequence of the hyperbolicity is,
539: \cite{Si68,Si77,Ga95c,GC95a,GC95b,Ga96},
540: that one can find a symbolic
541: representation of the points of $M$ in terms of sequences
542: $\underline{\varepsilon}=(\varepsilon_i)_{i=-\infty}^\infty$ of
543: finitely many digits $\varepsilon=1,\ldots,k$
544: subject only to a simple {\it hard core} restriction, namely
545: $T_{\varepsilon_j,\varepsilon_{j+1}}\equiv 1$
546: if $T$ is a matrix ({\it compatibility matrix})
547: with entries $0$ or $1$ and such that
548: $T^N_{\varepsilon,\varepsilon'}>0$ for some $N>0$
549: and all $\varepsilon,\varepsilon'$ ({\it mixing} condition). Moreover in such
550: a representation the dynamics becomes simply the left
551: shift, {\it i.e.} if $\underline{\varepsilon}(x)$
552: represents $x$ then $S(x)$ is represented by
553: the sequence $\underline{\varepsilon}$ shifted to the left by one unit.
554:
555: The key remarks are
556: \begin{enumerate}
557: \item Smooth observables on phase space can be
558: represented by {\it short range potentials}: in the case of the
559: observable $\sigma(x)$ this means that there are functions
560: $s_X(\underline{\varepsilon}_X)$
561: defined for all intervals $X=(a,\ldots,a+2n+1)$ and
562: $\underline{\varepsilon}_X=(\varepsilon_{a},
563: \ldots,\varepsilon_{a+2n+1})$, {\it translationally invariant}
564: $s_X=s_{X+b}$ and {\it exponentially decaying}
565: on time scale $\kappa^{-1}$ ({\it i.e.}
566: $|s_X(\underline{\varepsilon}_X)|< C e^{-\kappa
567: n}$ for some $C,\kappa>0$), such that
568: \beq
569: \label{FTC2.2}
570: \sigma(x)= s(\underline{\varepsilon}(x)),\qquad s(\underline{\varepsilon})=
571: \sum_{X\circ\, 0}
572: s_X(\underline{\varepsilon}_X)
573: \eeq
574: where the sum is over the intervals $X$ centered at the origin
575: (noted by $X\circ 0$).
576: Another important smooth observable is the {\it
577: expansion rate} $L(x)$ defined as the logarithm of the determinant
578: of the linearization matrix $\partial S(x)$ ({\it i.e.} the Jacobian
579: matrix of the map) restricted
580: to the unstable manifold: $L(x)=\log \det\partial S(x)_{u}$. This is also
581: expressible via an exponentially decaying potential $\Phi$:
582: \beq
583: \label{expansionrate}
584: L(x)=\ell(\underline{\varepsilon}(x)),\qquad
585: \ell(\underline{\varepsilon})=\sum_{X\circ\,0}
586: \Phi_X(\underline{\varepsilon}_X)
587: \eeq
588: \item The SRB distribution, represented as a distribution over
589: the (compatible) symbolic sequences $\underline{\varepsilon}$, is a
590: Gibbs state for the short
591: range potential $\Phi=(\Phi_X(\underline{\varepsilon}_X))$ defined in
592: Eq.~\ref{expansionrate}, {\it i.e.}
593: \beq
594: \label{15}
595: \langle F \rangle_{srb}=
596: \lim_{R\rightarrow\infty}
597: \frac{\sum_{\underline{\varepsilon}}e^{-\sum_{X\subset
598: \Lambda_R}\Phi_X(\underline{\varepsilon}_X)} F(\underline{\varepsilon}')}
599: {\sum_{\underline{\varepsilon}}e^{-\sum_{X\subset
600: \Lambda_R}\Phi_X(\underline{\varepsilon}_X)}}
601: \eeq
602: where $\Lambda_R=(-R,\ldots,R)\subset \ZZZ$, the sums extend over
603: compatible configurations $\underline{\varepsilon}=
604: (\varepsilon_{-R},\ldots,\varepsilon_R)$
605: ({\it i.e.} with
606: $T_{\varepsilon_j,\varepsilon_{j+1}}=1$ for $j=-R,\ldots,R-1$), and
607: $F(\underline{\varepsilon}')$ is an
608: arbitrary smooth observable defined on phase space regarded as a
609: function on the symbolic sequences and evaluated at a sequence
610: $\underline{\varepsilon}'$ which
611: extends (rather arbitrarily) $\underline{\varepsilon}$ to an infinite
612: compatible sequence by continuing $\V\e$ to the right with a sequence
613: $\V\e_{>}$
614: and to the left with a sequence
615: $\V\e_{<}$
616: into $\V\e'=(\V\e_{<},\V\e,\V\e_{>})$
617: so that $\V\e_{<}$ depends only on the symbol $\e_{-R}$ and
618: $\V\e_{>}$ depends only on the symbol $\e_{R}$: see
619: \cite{Si68,Ga02,GBG04}.
620: \end{enumerate}
621: The surprising reduction of the problem of studying the SRB
622: distribution to that of a {\it Gibbs distribution} for a one
623: dimensional chain with short range interaction (this is the physical
624: interpretation of Eq.~\ref{15}) generated the possibility of studying
625: more quantitatively at least some of the problems of nonequilibrium
626: statistical mechanics outside the domain of ``nonequilibrium
627: thermodynamics'', \cite{DGM}.
628:
629: \subsection{Finite time corrections to the characteristic function}
630:
631: The characteristic function $z_\tau(\lambda)$
632: of $p$, see Eq.~\ref{FTC2.1}, can therefore be
633: computed as
634: \beq
635: \label{4.1}
636: e^{-\tau z_\tau(\lambda)}=\lim_{R\rightarrow\infty}
637: \frac{\sum_{\underline{\varepsilon}}
638: e^{-\sum_{X\subset \Lambda_R}\Phi_X(\underline{\varepsilon}_X)
639: -\lambda \,\sum_{X\circ\, [0,\tau-1]}s_X(\underline{\varepsilon}_X) }}
640: {\sum_{\underline{\varepsilon}}
641: e^{-\sum_{X\subset\Lambda_R}\Phi_X(\underline{\varepsilon}_X)}}
642: \eeq
643: This means that it is the (limit as $R\rightarrow\infty$ of the) ratio between
644: the partition functions $Z_R(\Phi)$ of a Gibbs distribution in $\Lambda_R$
645: with potential $\Phi$
646: (the denominator)
647: and the partition function $Z_R(\Phi,\lambda s)$ with the same potential
648: {\it modified} in the {\it finite} region $[0,\tau-1]\subset\ZZZ$ by
649: the addition of a
650: potential $\lambda s_X(\underline{\varepsilon}_X)$.
651:
652: The one dimensional systems are very well understood and the above is a
653: well studied problem in statistical mechanics, known as {\it a
654: finite size corrections} calculation. For instance it can be attacked
655: by {\it cluster expansion}, \cite{GBG04}; this is a technique
656: to deal with the average of the exponential of a
657: spin Hamiltonian which is defined in terms of potentials $\phi_X$
658: exponentially decaying
659: with rate $\k$, such as those appearing in the numerator
660: and in the denominator of Eq.~\ref{4.1}. It allows us to represent
661: them as:
662: \beq \label{cluster}\sum_\ul\e e^{-\sum_{X\subset\L_R}\phi_X(\ul\e_X)}=
663: e^{-\sum_{X\subset\L_R}\wt\phi_X}\;,\eeq
664: where $\wt\phi_X$ are new {\it
665: effective} potentials, explicitly computable in terms of suitable averages
666: of products of $\phi_X(\ul\e_X)$'s, and which can be proven to be still
667: exponentially decaying with the diameter of $X$ with a rate $0<\k'\le\k$.
668:
669: In particular, a representation like Eq.~\ref{cluster}
670: allows us to rewrite the
671: partition function in the denominator of Eq.~\ref{4.1} as:
672: \beq
673: \label{4.2}
674: Z_R(\Phi)=\exp \left[ (2R+1) f_\infty(\Phi) - c_\infty(\Phi)+
675: O(e^{-\kappa'R}) \right]
676: \eeq
677: and the one in the numerator as
678: \beq
679: \begin{split}
680: \label{19}
681: &Z_R(\Phi,\lambda s)=\exp \Big[ (2R+1-\tau) f_\infty(\Phi)+
682: \tau f_\infty(\Phi+\lambda s) \\
683: &- c_\infty(\Phi)- g_\infty(\lambda)+ O(e^{-\kappa'R}+
684: e^{-\kappa'\tau}) \Big]
685: \end{split}
686: \eeq
687: Therefore
688: \beq
689: \begin{split}
690: \label{20}
691: z_\tau(\lambda)&=
692: f_\infty(\Phi)-f_\infty(\Phi+\lambda s)+
693: \frac{g_\infty(\lambda)}{\tau}+O(e^{-\kappa'\tau}) \\
694: &\defin z_\infty(\lambda)
695: +\frac{g_\infty(\lambda)}{\tau}+O(e^{-\kappa'\tau})
696: \end{split}
697: \eeq
698:
699: The function $z_\infty(\l)$ is
700: convex in $\l$ and the functions $g_\infty(\lambda)$ and
701: $z_\tau(\lambda)$ are analytic in $\lambda$ (a consequence of
702: the $1$-dimensionality and of the short range nature of the SRB
703: distribution): namely
704: $g_\infty(\lambda)=g^{(1)}_\infty \lambda
705: +\frac12g^{(2)}_\infty\lambda^2+\ldots$ and
706: $z_\tau(\lambda)=z^{(1)}_\tau\lambda
707: +\frac12z^{(2)}_\tau\lambda^2+\ldots$ and the coefficients of
708: their expansion in a power series of $\lambda$ can be expressed in
709: terms of correlation functions of $\sigma(x)$. For instance, from
710: Eq.~\ref{FTC2.1} and using the translational invariance of the SRB
711: measure, \beq\label{zio}
712: \begin{split}
713: z^{(1)}_\tau &= \tau^{-1} \langle \sum_{j=0}^{\tau-1}
714: \sigma(S^jx)\rangle_{srb} =
715: \sigma_+ \\
716: z^{(2)}_\tau &= \tau^{-1} \left[
717: \langle \sum_{j=0}^{\tau-1}\sigma(S^jx)\rangle_{srb}^2-
718: \langle \sum_{j=0}^{\tau-1} \sigma(S^jx)
719: \sum_{k=0}^{\tau-1}\sigma(S^kx) \rangle_{srb}
720: \right] \\
721: &=- \sum_{k=-\tau+1}^{\tau-1} \left[1-\frac{|k|}{\tau} \right]
722: \langle \sigma(S^kx) \sigma(x) \rangle_{c}
723: \end{split}
724: \eeq
725: where
726: $\langle \sigma(S^kx) \sigma(x) \rangle_{c} =
727: \langle \sigma(S^kx) \sigma(x) \rangle_{srb} - \sigma_+^2$.
728: Using Eq.~\ref{20},
729: $g_\infty(\lambda)=\lim_{\tau \rightarrow \infty}
730: \tau [ z_\tau(\lambda)-
731: z_\infty(\lambda) ]$, and the analyticity of
732: $z_\tau(\lambda)$, we have
733: $g^{(j)}_\infty=\lim_{\tau \rightarrow \infty}
734: \tau [ z^{(j)}_\tau - z^{(j)}_\infty ]$.
735: Since the connected correlation function
736: $\langle \sigma(S^kx) \sigma(x) \rangle_{c}$ {\it decays exponentially}
737: for $k \rightarrow \infty$, we obtain
738: \beq
739: \label{22}
740: \begin{split}
741: &g^{(1)}_\infty = 0 \\
742: &g^{(2)}_\infty = \sum_{k=-\infty}^\infty |k|
743: \langle \sigma(S^kx) \sigma(x) \rangle_{c}
744: \end{split}
745: \eeq
746:
747: \subsection{Finite time corrections to $\zeta_\infty(p)$}
748: \label{sec:IIC}
749:
750: A direct measurement of $z_\tau(\lambda)$ from
751: the numerical data is difficult. What is really accessible to
752: numerical observation are the quantities
753: $\frac1\tau\log \pi_\tau(\{p\in\Delta\})$ in Eq.~\ref{3} because the
754: measured values of $p$ are used to build an histogram obtained by
755: dividing the $p$--axis into sufficiently small
756: bins $\D$ and counting how many values of
757: $p$ fall in the various bins.
758: Let us choose the size of the bins $\D$ as
759: $|\D|=O(\e_\t/\t)$, with
760: $\e_\t$ a small parameter which will be eventually chosen $\e_\t=o(1)$,
761: see Appendix~\ref{app:A} for a discussion of this point.
762: Let also $p_\D$ be the center of
763: the bin $\D$. An application of a local form of central limit theorem,
764: discussed in Appendix~\ref{app:A}, shows that
765: the following asymptotic
766: representation of $\pi_\tau(\{p\in\Delta\})$ holds:
767: %
768: \beq\label{24}
769: \pi_\tau(\{p\in\Delta\})=e^{\t\z_\t(p_\D)}\Big(1+o(1)\Big)\eeq
770: %
771: where $\z_\t(p_\D)$ can be interpolated by an analytic
772: function of $p$, satisfying the equation
773: %
774: \beq\label{23}\z_\t(p)=-z_\t(\l_p)+\l_p p\s_+-\frac{1}{2\t}
775: \log\left[\frac{2\p}{\t}\Big(-\frac{z_\t''(\l_p)}{\s_+^2}\Big)\right]\eeq
776: %
777: and $\l_p$ is the inverse of $p(\l)=z_\t'(\l)/\s_+$.
778:
779: Using the previous equations, we now compute the
780: lowest order finite time correction to $\z_\io(p)$ around the maximum.
781:
782: We rewrite $\z_\t(p)$ as $\z_\t(p)=\z_\io(p)+\frac{\g_\io(p)}{\t}+O(\frac{1}
783: {\t^2})$.
784: By the analyticity of $\z_\t(p)$, we can
785: write $\z_\io(p),\g_\io(p)$ around $p=1$ in the form:
786: $\zeta_\infty(p) = \frac12\zeta_\infty^{(2)} (p-1)^2 +
787: \frac1{3!} \zeta_\io^{(3)} (p-1)^3 +
788: \ldots$ and
789: $\gamma_\infty(p) = \gamma_\infty^{(0)} + \gamma_\infty^{(1)} (p-1) +
790: \ldots$.
791:
792: Up to terms of order $(p-1)^2$ and
793: higher in the series for $\g_\io(p)$ we can rewrite:
794: %
795: \beq
796: \label{25}
797: \begin{split}
798: &\z_\t(p)=\\
799: &=\z_\io(p)+\frac{\g_\io^{(0)}}\t+
800: \frac{\g_\io^{(1)}}{\t}(p-1)
801: +O\left(
802: \frac{(p-1)^2}{\t}\right)+o\left(\frac1\t\right)=\\
803: &=\z_\io\left(p+
804: \frac{\g_\io^{(1)}}{\t\z^{(2)}_\io}\right)
805: +\frac{\g_\io^{(0)}}\t
806: +O\left(\frac{(p-1)^2}\t\right)+o\left(\frac1\t\right)\;.
807: \end{split}
808: \eeq
809: %
810: Thus, the finite time corrections
811: to $\z_\io(p)$ around its maximum begin with a shift of the maximum
812: at
813: %
814: \beq
815: \label{26}
816: p_0 = 1 - \frac{\gamma^{(1)}_\infty}{\tau \zeta^{(2)}_\infty} +
817: o\left(\frac1\tau\right)
818: \eeq
819: %
820: To apply the latter result we need to compute $\gamma^{(1)}_\infty$ in
821: terms of observable quantities. And, in order to compute $\gamma^{(1)}_\infty$
822: we apply Eq.~\ref{23}. First of all, we note that $\l_p$ is determined
823: by the condition
824: %
825: \beq \label{27} p\s_+=z'_\t(\l_p)=\s_++z_\t''(0)\l_p+O(\l_p^2)\eeq
826: %
827: where we used Eq.~\ref{zio} and Eq.~\ref{22}. Then,
828: $\l_p=\frac{\s_+(p-1)}{z_\t''(0)}+O\big((p-1)^2\big)$.
829: Substituting this result into Eq.~\ref{23} and equating the
830: terms of order $O(\frac{p-1}{\t})$ at both sides we find:
831: %
832: \beq \label{28}\g_\io^{(1)}=-\frac{1}{2}\frac{z_\io^{(3)}\s_+}{(z_\io^{(2)})^2}\;.
833: \eeq
834: %
835: The last equation can also be rewritten
836: as:
837: %
838: \beq\label{28a}
839: \gamma_\io^{(1)} = \frac{\zeta_\io^{(3)}}{2 \zeta_\io^{(2)}}
840: \eeq
841: %
842: This can be proven recalling that $\z_\io^{(2)}$ and
843: $\z_\io^{(3)}$
844: are derivatives of $\z_\io(p)$ in $p=1$, that
845: can be obtained by differentiating w.r.t. $\l$ (two or three times, respectively)
846: the definition
847: $\z_\io\big(z_\io'(\l)/\s_+\big)=-z_\io(\l)+\l z_\io'(\l)$ and computing the
848: derivatives in $\l=0$ recalling that $z_\io'(0)/\s_+=1$.
849: Plugging Eq.~\ref{28a} into Eq.~\ref{26} we finally get
850: %
851: \beq \label{wttau}p_0 = 1 - \frac{\z^{(3)}_\infty}{2\tau
852: (\zeta^{(2)}_\infty)^2} +
853: o\left(\frac1\tau\right)
854: \eeq
855: %
856: that is the main result of this section.
857: The key point is that
858: the moments $\z_\io^{(2)}$ and $\z_\io^{(3)}$ in Eq.~\ref{wttau} are quantities that
859: can be measured
860: from our empirical data (within an $O(\t^{-1})$ error).
861: We then have a verifiable prediction on the expected shift of the maximum
862: at finite $\t$. Our data agree very well with this prediction,
863: see Fig.~\ref{fig_2}
864: and corresponding discussion in sec.~\ref{sec:IV} below.
865:
866: Substituting Eq.~\ref{wttau} in Eq.~\ref{25},
867: we finally find:
868: %
869: \beq \label{oohh}
870: \z_\io(p)=
871: \h_\t(p)+O\Big(\frac{(p-1)^2}{\t}\Big)+o(\t^{-1})\;,\eeq
872: %
873: where $\h_\t(p)$ is defined as
874: %
875: \beq \label{etatau}
876: \h_\t(p)\defin =-\frac{\g^{(0)}_\io}{\t}+\z_\t\left(p-\frac{\z_\io^{(3)}}{
877: 2\t(\z_\io^{(2)})^2}\right)\;.\eeq
878: %
879: The key point of the above discussion was the validity of
880: Eq.~\ref{24}--\ref{23}; see Appendix~\ref{app:A} for their derivation.
881:
882: \subsection{Remarks}
883: \label{sec:IID}
884:
885: {\it (1)} The shift away from $1$
886: of the maximum of the function $\zeta_\tau(p)$ at
887: finite $\tau$, expressed by the second term in Eq.~\ref{etatau},
888: is due to the asymmetry of the distribution $\pi_\tau(p)$
889: around the average value $p=1$; consequently, it is proportional, at
890: leading order in $\tau^{-1}$, to $\zeta_\io^{(3)}$ which is
891: indeed a measure of the asymmetry of $\zeta_\io(p)$ around $p=1$.
892: This shift would be absent in the case of a symmetric distribution
893: ({\it e.g.}, a Gaussian) and for this reason it was not observed in
894: previous experiments \cite{ECM93,BGG97,BCL98,ZRA04}.
895:
896: {\it (2)} The error term in the r.h.s. of Eq.~\ref{24} is $o(1)$
897: w.r.t. $\t$ and it does not affect the computation of
898: $\g_\io(p)$. It is then clear that with a calculation similar to that
899: we performed, one can get equations for the coefficients $O(\l^k)$ in the
900: exponents of Eq.~\ref{24}; in this way
901: one can iteratively construct the whole sequence of coefficients
902: $\g_\io^{(k)}$ defining the power series expansion of $\g_\io(p)$.
903:
904: {\it (3)} In models with continuous time evolution the quantity $\sigma_+$ is
905: not dimensionless but it has dimensions of inverse time: in such cases
906: one can imagine that one is still studying a map which maps a system
907: configuration at a time when some prefixed event happens in the system
908: (typically a ``collision'') into the next one in which a similar event
909: takes place. If $\tau_0$ is the average time interval between such events
910: then $\tau_0\sigma_+$ will play the role played by $\sigma_+$ in the
911: discrete time case: it will be the adimensional parameter entering
912: the estimates of the error terms.
913:
914: Note that the coefficients
915: $g_\io^{(k)}$ are of order $\s_+^k$, and their
916: size is necessarily estimated by (the adimensional) entropy production to
917: the $k$--th power. Then, in the continuous time case, the choice
918: of $\t_0$ affects the estimates of the remainders, because it
919: affects the size of the adimensional parameter $\t_0\s_+$;
920: and the size of the mixing time (that is connected with the estimated range of
921: decay of the potentials, see \cite{GBG04}). The natural (and physical)
922: choice for $\t_0$ is the mixing time. Consistently with this remark,
923: at the moment of constructing numerically the
924: distribution function for the entropy
925: production rate averaged over a time $\t$, we will always
926: consider time intervals of the form
927: $\t=\t_0 n$, $n\ge 1$, see section~\ref{sec:IIIC} below.
928:
929: \section{The model}
930: \label{sec:III}
931:
932: We consider a system of $N$ classical particles of equal mass $m$
933: in dimension $d$; they are described by their position $q_i$ and momenta
934: $p_i = m \dot q_i$,
935: $(p_i,q_i) \in R^{2d}, \ i=1, \ldots, N$.
936: The particles are confined in a cubic box of side $L$ with periodic
937: boundary conditions.
938: Each particle is subject to a {\it conservative force},
939: $f_i(\ul q) = - \partial_{q_i} V(\ul q)$, and to a
940: {\it nonconservative force} $E_i$ that does not depend on
941: the phase space variables. The force $E_i$ is locally conservative
942: but not globally such due to periodic boundary conditions.
943: The {\it mechanical thermostat} is a Gaussian thermostat \cite{EM90},
944: $\theta_i (\ul p,\ul q) = -\alpha(\ul p, \ul q) \, p_i$, and the
945: function $\alpha(\ul p, \ul q)$ is defined by the condition that the
946: total kinetic energy
947: $K(\ul p) \equiv \frac1{2m} |\ul p|^2 = \frac1{2m} \sum_i p^2_i$
948: should be a constant
949: ({\it isokinetic ensemble}).
950: The equations of motion are:
951: \beq
952: \label{eqofmotion}
953: \begin{cases}
954: &\dot{q}_i = \frac{p_i}{m} \ , \\
955: &\dot{p}_i = f_i(\ul q) + E_i - \alpha(\ul p,\ul q) \ p_i \ ,
956: \end{cases}
957: \eeq
958: From the constraint $\frac{dK}{dt} = 0$ one obtains
959: \beq
960: \label{alphadef}
961: \alpha(\ul p,\ul q)=\frac{\sum_i E_i \ p_i +
962: \sum_i f_i(\ul q) \ p_i}{\sum_i p^2_i} \ .
963: \eeq
964:
965: \subsection{Entropy production rate}
966:
967: The total phase space volume contraction rate for this system is given by:
968: \beq
969: \begin{split}
970: &\sigma(\ul p,\ul q)=-\sum_i \left( \frac{\partial \dot{q}_i}{\partial q_i}
971: +\frac{\partial \dot{p}_i}{\partial p_i} \right) \\
972: &= dN\alpha(\ul p,\ul q) + \sum_i \frac{\partial \alpha}{\partial p_i} p_i
973: = (dN-1) \ \alpha(\ul p,\ul q) \ .
974: \end{split}
975: \eeq
976: Defining the {\it kinetic temperature}, $T \equiv 2 K(\ul p)/(dN-1)$,
977: \cite{EM90}, the phase space contraction rate can be rewritten
978: as
979: \beq\label{36zz}
980: \sigma(\ul p, \ul q) =
981: \frac{\sum_i E_i \ \dot q_i - \dot V}{T} \ .
982: \eeq
983: The first term is the power dissipated by
984: the external force divided by the kinetic
985: temperature, and can be identified with the entropy production rate
986: \cite{EM90,ECM93,Ga04}.
987: The second term is the total derivative w.r.t. time of the potential energy
988: divided by the temperature: this term does not affect the validity of the
989: Fluctuation Relation in the asymptotic limit $\t\to\io$, as total derivatives
990: give a contribution $O(\tau^{-1})$ in $p(x)$ \cite{Ga04,Ru00}, hence
991: they do not contribute to $\zeta_\io$; however it has effect on the
992: distribution of fluctuations over a finite time $\t$ and
993: its influence on the
994: numerical computations has been recently discussed in detail \cite{ZRA04}.
995: The most convenient thing to do, in order to have a finite time
996: distribution that
997: approximates in the best possible way the asymptotic distribution
998: of fluctuations,
999: is to study the distribution of fluctuations for
1000: the {\it entropy production rate} $\dot s$, where $\dot s$ is identified with
1001: $\s$ {\it minus} the total derivative term $-\dot V/T$ in Eq.~\ref{36zz}:
1002: %
1003: \beq \label{epr}\dot s(\ul p, \ul q) =
1004: \frac{\sum_i E_i \ \dot q_i}{T}
1005: \eeq
1006: %
1007: From now on we will call $\z_\io(p)$ and $\z_\t(p)$ the distributions
1008: for the fluctuations of the entropy production rate $\dot s$ averaged
1009: over infinite or finite time, respectively. These will be the
1010: objects we will measure and use from now on.
1011:
1012: In order to define the {\it current} $J(x,E)$,
1013: let us rewrite $E_i = E \, u_i$, where $u_i$
1014: is a (constant) unit vector that specifies
1015: the direction of the force acting on the $i$-th particle.
1016: Then, according to Eq.~\ref{Jdef},
1017: \beq
1018: \label{Jcolor}
1019: J(\ul p, \ul q) =
1020: \frac{\partial \sigma}{\partial E} = \frac{\sum_i u_i \, \dot q_i}{T}
1021: \eeq
1022:
1023: \subsection{Discretization of the equations of motion}
1024:
1025: To perform the numerical simulation, one has to write the
1026: equations of motion in a discrete form.
1027: One possibility is to use the {\it Verlet algorithm} \cite{AT87};
1028: for Hamiltonian equations of motion ({\it i.e.}, $\ul E=\ul 0$ and $\alpha=0$)
1029: \beq
1030: \label{simpleeqofmotion}
1031: \begin{cases}
1032: &\dot{q}_i = \frac{p_i}{m} \ , \\
1033: &\dot{p}_i = f_i(\ul q) \ ,
1034: \end{cases}
1035: \eeq
1036: the Verlet discretization has the form
1037: \beq
1038: \begin{cases}
1039: &q_i(t+dt) = q_i(t) + \frac{p_i(t)}{m} dt + \frac12 f_i(t) dt^2 \ , \\
1040: &p_i(t+dt) = p_i(t) + \frac12 \big[ f_i(t) + f_i(t+dt) \big] dt \ ,
1041: \end{cases}
1042: \eeq
1043: where $dt$ is the {\it time step size}. This discretization ensures that
1044: the error is $O(dt^4)$ on the positions $q_i(t)$ in a single time step.
1045: The implementation of this algorithm on a computer is discussed in
1046: detail in \cite{AT87}.
1047:
1048: However, this method requires the forces $f_i(t)$ to depend only on the
1049: positions and not on the velocities: hence, it has to be adapted
1050: to Eq.s~\ref{eqofmotion}.
1051: This has been done in the following way.
1052: We write the discretized equations as
1053: \beq
1054: \label{eqdiscrete}
1055: \begin{cases}
1056: q_i(t+dt) &= q_i(t) + \frac{p_i(t)}{m} dt \\
1057: &+ \frac12 \big[ f_i(t) + E_i - \alpha(t) p_i(t) \big] dt^2 \ , \\
1058: p_i(t+dt) &= p_i(t) + E_i +\frac12 \big[ f_i(t) + f_i(t+dt) \\
1059: &- \alpha(t) p_i(t) - \alpha(t+dt) p_i(t+dt) \big] dt \ ,
1060: \end{cases}
1061: \eeq
1062: with the same error as in the standard Verlet discretization.
1063: We store in the computer, at time $t$,
1064: the positions $q_i(t)$, the momenta $p_i(t)$,
1065: the forces $f_i(t)$, and the Gaussian multiplier $\alpha(t)$.
1066: Then, we perform the following operations:
1067: \begin{enumerate}
1068: \item we calculate the new positions $q_i(t+dt)$ using the first equation;
1069: \item using the new positions we calculate the new forces $f_i(t+dt)$ (the
1070: conservative forces depend only on the positions);
1071: \item we calculate the quantity
1072: $\xi_i = p_i(t) + E_i +\frac12 \big[ f_i(t) + f_i(t+dt)
1073: - \alpha(t) p_i(t) \big] dt$ and we observe that
1074: $p_i(t+dt)$ can be expressed in terms of the (known) $\x_i$ and the (unknown)
1075: $\a(t+dt)$ as
1076: \beq
1077: \label{DISCRE1}
1078: p_i(t+dt)=\frac{\xi_i}{1 - \alpha(t+dt) dt/2} \ ;
1079: \eeq
1080:
1081: \item substituting Eq.~\ref{DISCRE1} in the definition of $\alpha(t+dt)$,
1082: Eq.~\ref{alphadef},
1083: we get a self-consistency equation for $\alpha(t+dt)$, whose solution is
1084: \beq
1085: \label{selfconsist}
1086: \begin{split}
1087: &\alpha(t+dt) = \frac{\alpha_0}{1-\alpha_0 dt /2} \ , \\
1088: &\alpha_0 = \frac{ \sum_i E_i \ \xi_i + \sum_i f_i(t+dt)
1089: \ \xi_i }{\sum_i \xi_i^2} \ ;
1090: \end{split}
1091: \eeq
1092: \item substituting Eq.~\ref{selfconsist} in Eq.~\ref{DISCRE1}
1093: we calculate $p_i(t+dt)$.
1094: \end{enumerate}
1095: This procedure allows us to calculate the new positions,
1096: momenta, forces, and $\alpha$,
1097: at time $t+dt$ according to Eq.s~\ref{eqdiscrete}
1098: {\it without approximations},
1099: defining a map $S$ such that $(\ul p(t+dt),\ul q(t+dt))= S(\ul p(t),\ul q(t))$.
1100:
1101: Our ({\it discrete}) dynamical system
1102: will be defined by the map $S(\ul p,\ul q)$
1103: and will approximate the differential equations of motion,
1104: Eq.~\ref{eqofmotion}, with error $O(dt^4)$
1105: for the positions and $O(dt^3)$ for the velocities. \\
1106: The map $S$ satisfies the following properties:
1107: \begin{enumerate}
1108: \item it is {\it reversible}, {\it i.e.} it
1109: exists a map $I(\ul p, \ul q)$ (simply
1110: defined by $I(\ul p, \ul q)=(-\ul p, \ul q)$)
1111: such that $I S = S^{-1} I$;
1112: \item in the {\it Hamiltonian} case ($\ul E=\ul 0$ and $\alpha=0$,
1113: Eq.s~\ref{simpleeqofmotion}) it is {\it volume preserving}.
1114: \end{enumerate}
1115: The first property ensures that {\it assuming
1116: the Chaotic Hypothesis} the Fluctuation Relation holds for the map $S$.
1117: The second property ensures that at equilibrium the discretization
1118: algorithm conserves the phase space volume.
1119:
1120: \subsection{Details of the simulation}
1121: \label{sec:IIIC}
1122:
1123: In the simulation, we chose the external force of the form
1124: $E_i = E \, u_i$, where
1125: the unit vectors $u_i$ were parallel to the $x$
1126: direction but with different orientation: half of them were oriented in
1127: the positive
1128: direction, and half in the negative direction, {\it i.e.} $u_i = (-1)^i \wh x$,
1129: in order to keep the center of mass fixed.
1130: We considered two different systems, selecting interaction potentials
1131: widely used in numerical simulations (for the purpose
1132: of making easier possible future independent checks and rederivations of our
1133: results):
1134: \begin{enumerate}
1135: \item ({\it model I}) the first investigated system is made by $N=8$
1136: particles of equal mass $m$ in $d=2$.
1137: The interaction potential is a sum of pair interactions,
1138: $V(\ul q)= \sum_{i < j} v(|q_i - q_j|)$,
1139: and the pair interaction is represented by a
1140: WCA potential, {\it i.e.} a
1141: Lennard-Jones potential truncated at the minimum:
1142: \beq
1143: \nonumber
1144: v(r) =
1145: \begin{cases}
1146: & 4 \epsilon \left[ \left(\frac\sigma r \right)^{12} -
1147: \left(\frac\sigma r \right)^6 \right] + \epsilon \ ,
1148: \hspace{10pt} r \leq \sqrt[6]{2}\sigma \ ; \\
1149: & 0 \ , \hspace{100pt} r > \sqrt[6]{2}\sigma \ .
1150: \end{cases}
1151: \eeq
1152: The reduced density was $\rho=N\sigma^2/L^2=0.95$
1153: (that determines $L$),
1154: the kinetic temperature was fixed to $T=4\epsilon$
1155: and the {\it time step} to $dt=0.001 t_0$, where
1156: $t_0 = \sqrt{m\sigma^2/\epsilon}$.
1157: In the following, all the quantities will be
1158: reported in units of $m$, $\epsilon$
1159: and $\sigma$ ({\it LJ units}).
1160: This system was already studied in the literature,
1161: see {\it e.g.} \cite{ECM93,SEC98}.
1162: We investigated different values of the external force $E$ ranging from
1163: $E=0$ to $E=25$.
1164: \item ({\it model II})
1165: the second system is a binary mixture of $N$=20 particles (16 of type A and
1166: 4 of type B), of equal mass
1167: $m$, in $d=3$, interacting via the same WCA potential of model I; the pair
1168: potential is
1169: \beq
1170: \nonumber
1171: v_{\alpha\beta}(r) =
1172: \begin{cases}
1173: & 4 \epsilon_{\alpha\beta} \left[
1174: \left(\frac{\sigma_{\alpha\beta}} r \right)^{12}
1175: - \left(\frac{\sigma_{\alpha\beta}} r \right)^{6} \right]
1176: + \epsilon_{\alpha\beta} \ ,\\
1177: & \hspace{100pt} r \leq \sqrt[6]{2}\sigma_{\alpha\beta} \ ; \\
1178: & 0 \ , \hspace{88pt} r > \sqrt[6]{2}\sigma_{\alpha\beta} \ ;
1179: \end{cases}
1180: \eeq
1181: $\alpha$ and $\beta$ are indexes that specify the particle species
1182: ($\alpha,\beta \in \{A,B\}$).
1183: The parameters entering the potential are the following:
1184: $\sigma_{AB}=0.8 \sigma_{AA}$;
1185: $\sigma_{BB}=0.88 \sigma_{AA}$;
1186: $\epsilon_{AB}=1.5 \epsilon_{AA}$;
1187: $\epsilon_{BB}=0.5 \epsilon_{AA}$.
1188: Similar potentials have been studied, \cite{KA94,DmSC04},
1189: as models for liquids in the
1190: supercooled regime ({\it i.e.}, below the melting
1191: temperature).
1192: For this system the {\it LJ units} are
1193: $m$, $\epsilon_{AA}$, and $\sigma_{AA}$; the unit of time
1194: is then $t_0 = \sqrt{ m\sigma_{AA}^2/\epsilon_{AA}}$.
1195: The reduced density was $\rho = N \sigma_{AA}^3 / L^3 = 1.2$
1196: and the integration step was $dt=0.001 t_0$.
1197: The unit vectors $u_i$ are chosen such that half of the $A$
1198: particles and half of the
1199: $B$ particles have positive force in the $x$ direction,
1200: and the remaining particles
1201: have negative force in the $x$ direction.
1202: For this system we investigated different values of external
1203: force $E \in [0,10]$ and temperature $T \in [0.5,3]$.
1204: \end{enumerate}
1205:
1206: For each system and for each chosen value of $T$ and $E$, we simulated a very
1207: long trajectory ($\sim 2 \cdot 10^9 dt$)
1208: starting from a random initial data; we recall that in both
1209: systems we chose $dt=
1210: 0.001 t_0$, $t_0$ being the natural unit time introduced in items (1)
1211: and (2) above. After a
1212: short transient ($\sim 10^3 dt$), still much bigger than the
1213: decay time $\t_0$ of self-correlations (that
1214: appears to be $\t_0= 10^2 dt$),
1215: the system reached stationarity, in the sense that the instantaneous
1216: values of observables (\eg potential energy, Lyapunov exponents)
1217: agree with the corresponding asymptotic values within the statistical error
1218: of the asymptotic values themselves.
1219: After this transient we started recording
1220: values $p_i$, $i=1,\ldots,{\cal N}$,
1221: of the variable $p(x)$ (defined in Eq.~\ref{pdix}),
1222: integrating the entropy production rate (Eq.~\ref{epr})
1223: on adjacent segments of trajectory of length $\tau_0 = 100 dt = 0.1 t_0$.
1224: Note that the length of the time interval over which we averaged the entropy
1225: production rate was chosen as equal to the mixing time, consistently
1226: with the discussion in Remark~(4) of sec.~\ref{sec:IID}.
1227:
1228: In conclusion, from each simulation run at fixed $T$ and $E$
1229: we obtain ${\cal N} \sim 10^7$ values $p_i$ of $p(x)$ which are
1230: the starting point of our data analysis.
1231: The value of $\sigma_+$ is estimated by
1232: averaging the entropy production rate over the whole trajectory.
1233:
1234: From a shorter simulation run we measured also the Lyapunov exponents
1235: of the map $S$ using the standard algorithm of Benettin {\it et al.}
1236: \cite{SEC98,BGS76}.
1237:
1238: \subsection{Remarks}
1239:
1240: To conclude this section, we note that the WCA potential has a discontinuity in
1241: the second derivative. Thus, one should be concerned with the possibility
1242: that the error of our discretization is not $O(dt^4)$ over the $q_i$'s
1243: on a single time step, as it should be for potentials $V \in C^4$.
1244: To check that this is not the case (or that
1245: at least this does not affect our results) we made two independent tests:
1246: \begin{enumerate}
1247: \item we simulated a system similar to model I but with a potential $V \in C^4$
1248: and we obtained qualitatively the same results;
1249: \item we simulated model I using an
1250: {\it adaptive step size} algorithm \cite{AT87};
1251: this kind of algorithms adapt the step size $dt$ during the simulation
1252: in order to keep constant the difference
1253: between a single step of size $dt$ and two
1254: steps of size $dt/2$. If the precision of our
1255: discretization changed at the singular
1256: points of the potential, the time step should change
1257: abruptly during the simulation,
1258: while we observed a practically constant time step during the simulation.
1259: \end{enumerate}
1260: Hence, we have evidence of the fact that the (isolated) singularities
1261: of the potentials do not produce
1262: relevant effects on our observations;
1263: this is probably due to the fact that the set of singular points of the total
1264: potential energy $V(\ul q)$ has zero measure w.r.t. the SRB measure.
1265:
1266: \section{Data analysis}
1267: \label{sec:IV}
1268:
1269: \begin{figure}[t]
1270: \centering
1271: \includegraphics[width=.50\textwidth,angle=0]{FIG1.eps}
1272: \caption{Model I at $E=5$: the function
1273: $\eta_\tau(p)=\zeta_\infty(p)+O((p-1)^2/\tau)$
1274: for different
1275: values of $\tau$.}
1276: \label{fig_1}
1277: \end{figure}
1278:
1279: In this section we will discuss in detail the procedure we followed to
1280: analyze the numerical data. As an example, we will discuss the
1281: data obtained from the simulation of model I at $E=5$. As discussed in the
1282: previous section, from the simulation run we obtain a set
1283: ${\cal P}_0 = \{ p_i \}_{i=1\ldots{\cal N}}$ of values of the variable $p(x)$
1284: that correspond to $\tau= \tau_0$ and are
1285: measured on adjacent segments of trajectory.
1286: We recall again that $\tau_0 = 0.1 = 100 dt$ is of the order
1287: of the {\it mixing time}, {\it i.e.}
1288: the time scale over which the correlation functions
1289: ({\it e.g.} of density fluctuations) decay to zero.
1290:
1291: \vskip10pt
1292:
1293: \noindent
1294: {\it Probability distribution function} -
1295: From the dataset ${\cal P}_0$ we construct the histograms
1296: $\pi_\tau(p)$ for different values of $\tau = n \tau_0$ as follows:
1297: the values of $p(x)$ for $\tau = n\tau_0$ are obtained by averaging
1298: $n$ subsequent entries of the dataset ${\cal P}_0$; we obtain a new dataset
1299: ${\cal P}_n = \{ p^{(n)}_j \}_{j=1\ldots{\cal N}/n}$ such that
1300: $p^{(n)}_j= n^{-1} \sum_{i=nj+1}^{n(j+1)} p_i$.
1301: Finally, from the dataset ${\cal P}_n$ the histogram of $\pi_{\tau}(p)$ is
1302: constructed for $\tau=n\tau_0$;
1303: the errors are estimated as the square roots of the number of counts
1304: in each bin.
1305: The function $\zeta_{\tau}(p)$ is then defined as
1306: $\zeta_\t(p)=\t^{-1} \log \pi_\t(p)$.
1307:
1308: \begin{figure}[t]
1309: \includegraphics[width=.50\textwidth,angle=0]{FIG2.eps}
1310: \caption{Model I at $E=5$: the maximum $\widetilde p_\tau$ of $\zeta_\tau(p)$
1311: as a function of $1/\tau$.
1312: The full line is the prediction of Eq.~\ref{wttau},
1313: $\widetilde p = 1 - \zeta_\io^{(3)}/\big[2\tau
1314: \big(\zeta_\io^{(2)}\big)^2\big]$.}
1315: \label{fig_2}
1316: \end{figure}
1317:
1318: \vskip10pt
1319:
1320: \noindent
1321: {\it Shifting of the maximum} -
1322: By fitting the function $\zeta_{\tau}(p)$ in $p\in[-1,3]$
1323: with a sixth-order polynomial we determine the position of the maximum
1324: $\widetilde p_\tau$ within an error that, since $\d p$ is the length of a bin,
1325: we estimate to be $\d p/2$.
1326: Then, we construct the function
1327: $\eta_{\tau}(p) = \zeta_\tau(p-1+\widetilde p_\tau)$
1328: (see Eq.~\ref{etatau}) which
1329: is expected to approximate the limiting function $\zeta_\io(p)$ with error
1330: $O((p-1)^2/\tau)$.
1331: The functions $\eta_\tau(p)$ are reported
1332: in Fig.~\ref{fig_1} for different
1333: values of $\tau$.
1334: We observe a very good convergence for $\tau \gtrsim 5.0 = 50\tau_0$.
1335:
1336: By a fourth-order fit of the so-obtained limiting function $\zeta_\io(p)$
1337: around $p=1$ we extract the coefficients $\zeta_\io^{(2)}=-0.287$ and
1338: $\zeta_\io^{(3)}=0.149$ in order to test the correctness of
1339: Eq.~\ref{wttau}.
1340: In Fig.~\ref{fig_2} we report $\widetilde p_\tau$.
1341: The full line is the prediction of
1342: Eq.~\ref{wttau}, that is indeed verified for $\tau \gtrsim 10$.
1343: This result confirms the analysis of section~\ref{sec:II}.
1344:
1345: \begin{figure}[t]
1346: \includegraphics[width=.50\textwidth,angle=0]{FIG3.eps}
1347: \caption{Model I at $E=5$:
1348: the estimate of the function $\zeta_\io(p)$ (open circles).
1349: In the same
1350: plot $\zeta_\io(-p) + p\sigma_+$ (filled squares)
1351: is reported. In the inset, the
1352: interval $p\in[-2,2]$ where the data overlap
1353: is magnified. The full
1354: line is the Gaussian approximation,
1355: $\frac12\zeta_\io^{(2)}(p-1)^2$.
1356: The plot shows that the Gaussian is not a good approximation
1357: in the interval $[-2,2]$. The validity of the Fluctuation Relation
1358: in the same interval is shown by the overlap of the open circles and
1359: filled squares.}
1360: \label{fig_3}
1361: \end{figure}
1362:
1363: \vskip10pt
1364:
1365: \noindent
1366: {\it Graphical verification of the fluctuation relation} -
1367: From the previous analysis we can conclude that the function $\eta_\tau(p)$
1368: for $\tau = 5.0$ provides a good estimate
1369: of the function $\zeta_\io(p)$ for $p\in[-2,4]$
1370: (see Fig.~\ref{fig_1});
1371: thus, we can use this function to test the
1372: fluctuation relation, Eq.~\ref{4},
1373: in this range of $p$. In Fig.~\ref{fig_3} we report the estimated functions
1374: $\zeta_\io(p)$ and $\zeta_\io(-p)+p\sigma_+$. An excellent agreement between
1375: the two functions is observed in the interval $p\in[-2,2]$ where our data
1376: allows the computation of both $\zeta_\io(p)$ and $\zeta_\io(-p)$.
1377: Note that in this range of $p$ the function $\zeta_\io(p)$ is not Gaussian, see
1378: the inset of Fig.~\ref{fig_3}.
1379:
1380: \vskip10pt
1381: \noindent
1382: {\it Quantitative verification of the fluctuation relation} -
1383: The translation of the function $\zeta_\tau(p)$ is crucial to obtain a
1384: correct estimate of the limit $\zeta_\infty(p)$ and to verify the fluctuation
1385: relation.
1386: In this section we will try to quantify this observation;
1387: as the discussion will
1388: be very technical, the reader who is satisfied with Fig.~\ref{fig_3} should
1389: skip to next section.
1390:
1391: The histogram $\pi_{n\tau_0}(p)$ derived from the
1392: dataset ${\cal P}_n$ is constructed
1393: assigning the number of counts $\pi_\alpha$ in the
1394: $\alpha$-th bin to the middle
1395: of the binning interval, that we call $p_\alpha$
1396: (the latter will be an {\it increasing} function of $\alpha$).
1397: The statistical error
1398: $\delta \pi_\alpha$ on the number of counts is $\sqrt{\pi_\alpha}$.
1399: Our histograms are constructed
1400: in such a way that if $p_\alpha$ is the center
1401: of a bin, also $-p_\alpha$ is the center
1402: of a bin; we call $\overline\alpha$ the bin
1403: such that $p_{\overline\alpha}=-p_\alpha$.
1404: There exists a value $p_m$ such that for $p_\alpha < p_m$ the
1405: number of counts in the bin $\alpha$ is smaller than $m$ (we choose $m=4$).
1406: Let us indicate by $p_{\alpha_m}$ the smallest value of $p_\alpha > p_m$.
1407: Hence, the histogram is characterized by:
1408: \begin{enumerate}
1409: \item a {\it bin size} $\delta p$;
1410: \item
1411: the bin $\alpha_m$ corresponding to the minimum value of $p_\alpha$ such
1412: that the number of counts in the bin is at least $m$;
1413: \item the total number $M$ of bins such that
1414: $\alpha \in [\alpha_m,\overline\alpha_m]$; for these values of $p_\alpha$,
1415: both $\pi_\tau(p)$ and $\pi_\tau(-p)$
1416: can be computed and they can be used to verify
1417: the fluctuation relation.
1418: \end{enumerate}
1419: The function $\zeta_{\tau}(p)$, derived from the histogram, is specified by
1420: a set of values $(p_\alpha,\zeta_\alpha,\delta \zeta_\alpha)$
1421: for each bin $\alpha$,
1422: where $\zeta_\alpha=\tau^{-1} \log \pi_\alpha$ and
1423: the error $\delta\zeta_\alpha$ has been defined by
1424: \beq
1425: \delta \zeta_\alpha = \frac1\tau \frac{\delta \pi_\alpha}{\pi_\alpha}=
1426: \frac1{\tau \sqrt{\pi_\alpha}} \ .
1427: \eeq
1428: A quantitative verification of Eq.~\ref{4}
1429: is possible defining the following
1430: $\chi^2$ function:
1431: \beq
1432: \chi^2 \equiv \frac1M \sum_{\alpha=\alpha_m}^{\overline\alpha_m}
1433: \frac{(\zeta_\alpha - \zeta_{\overline\alpha} - p_\alpha \sigma_+)^2}
1434: {(\delta \zeta_\alpha)^2 + (\delta \zeta_{\overline\alpha})^2}
1435: \eeq
1436: The value of $\chi$ is the average difference between $\zeta_\tau(p)$ and
1437: $\zeta_\tau(-p)+p\sigma_+$ in units of the statistical error.
1438: Translating $p$ of a quantity $a \delta p/2$, $a\in \ZZZ$,
1439: corresponds to shifting the histogram, {\it i.e.} to
1440: consider a new histogram $(p_\alpha + a \delta p/2,\zeta_\alpha,
1441: \delta\zeta_\alpha)$.
1442: This preserves the property that if $p_\alpha$ is the center of a bin, also
1443: $-p_\alpha$ is the center of a bin; we
1444: call $\overline\alpha(a)$ the new value of $\alpha$
1445: such that $p_{\overline\alpha(a)}+a \delta p/2=
1446: -(p_\alpha + a \delta p/2)$. Also, the number $M_a$ of bins such that
1447: $\alpha(a) \in [\alpha_m, \overline\alpha_m(a)]$ depends on $a$.
1448: We define
1449: \beq
1450: \label{chia}
1451: \chi^2(a) \equiv \frac1{M_a} \sum_{\alpha=\alpha_m}^{\overline\alpha_m(a)}
1452: \frac{\big(\zeta_\alpha - \zeta_{\overline\alpha(a)} -
1453: (p_\alpha + a\delta p/2) \sigma_+\big)^2}
1454: {(\delta \zeta_\alpha)^2 + (\delta \zeta_{\overline\alpha(a)})^2}
1455: \eeq
1456: We shall use the criterion that
1457: the fluctuation relation is satisfied if $\chi \leq 3$, which means that
1458: $\zeta_\io(p)$ and $\zeta_\io(-p)+p \sigma_+$
1459: differ, {\it on average}, by less than $3$ times the statistical error
1460: $\sqrt{ \big( \delta \zeta(p)\big)^2 +\big( \delta \zeta(-p)\big)^2}$.
1461: The function $\chi(a)$ for the case of model I at $E=5$
1462: is reported in Fig.~\ref{fig_4}.
1463:
1464: \begin{figure}[t]
1465: \includegraphics[width=.33\textwidth,angle=0]{FIG4.eps}
1466: \caption{Model I at $E=5$:
1467: the function $\chi(a)$. The full line corresponds to $\chi=3$.
1468: The arrow indicates the interval $\d_0\pm\d p/2$
1469: (note that its length is 2 in units of $a$)
1470: into which the minimum of $\chi$ can be located
1471: within the accuracy of the histogram.}
1472: \label{fig_4}
1473: \end{figure}
1474:
1475: The minimum of $\chi$
1476: is assumed between $a^*=1$ and $a^*+1=2$ and an upper limit for the value of
1477: $\chi$ at the minimum is $\chi(1)=3.5$.
1478: We estimate the translation that minimizes $\chi$ as
1479: $\delta_0 = (a^*+0.5) \delta p/2 = 1.5 \cdot 0.093 = 0.140$,
1480: and to this estimate
1481: we attribute an error $\pm\d p/2$, where $\delta p =
1482: 0.186$ is the size of a bin.
1483: On the other hand, we have seen above that,
1484: in order to shift the maximum of $\zeta_\tau(p)$
1485: in $p=1$, one has to translate $p$ by a
1486: quantity $\delta \equiv 1 - \widetilde p = 0.215$.
1487: The consistency of our analysis
1488: requires that $\delta$ and $\d_0$ coincide within their errors, \ie
1489: that the intervals $\d\pm\d p/2$ and $\d_0\pm\d p/2$ overlap, or in other
1490: words $|\d-\d_0|<\d p$. In the present case $0.075=|\d-\d_0|<\d p=0.186$,
1491: then $\d$ and $\d_0$ coincide within the errors. This means that
1492: the translation of $p$
1493: brings the maximum of $\zeta_\tau(p)$ in
1494: $p=1$ and, {\it at the same time}, minimizes the difference between
1495: $\h_\tau(p)$ and $\h_\tau(-p)+p \sigma_+$, where $\h_\t$ is our
1496: finite time estimate of $\z_\io(p)$.
1497: The value $\chi(a^*)$ quantifies
1498: this difference and is a first estimate of the precision of our analysis.
1499:
1500: \begin{figure}[t]
1501: \includegraphics[width=.33\textwidth,angle=0]{FIG5.eps}
1502: \caption{Model I at $E=5$: the function $\overline X(a)$.
1503: The horizontal arrow marks the interval where
1504: the minimum of $\chi$ is located, see
1505: Fig.~\ref{fig_4}.
1506: The vertical arrow indicates the error
1507: $\delta X$ on the value $X=1$ which is estimated
1508: as $\delta X=2(\overline X(2)-\overline X(1))$.
1509: The optimal slope of the fluctuation relation without
1510: the translation would have been $\overline X(a=0) \sim 0.85$.}
1511: \label{fig_5}
1512: \end{figure}
1513:
1514: Another estimate of the precision of our analysis can be obtained as follows.
1515: We define a parameter $X$ as the slope of $\zeta_\io(p)-\zeta_\io(-p)$ as a
1516: function of $p\sigma_+$:
1517: \beq
1518: \zeta_\io(p)= \zeta_\io(-p)+Xp\sigma_+
1519: \eeq
1520: The fluctuation theorem predicts $X=1$, but other values of $X$ are possible
1521: under different hypothesis, see \cite{BG97,BGG97,Ga04,Ga99b}.
1522: We define a function $\chi^2(a,X)$ as
1523: \beq
1524: \label{chiaX}
1525: \chi^2(a,X) \equiv \frac1{M_a}
1526: \sum_{\alpha=\alpha_m}^{\overline\alpha_m(a)}
1527: \frac{\big(\zeta_\alpha - \zeta_{\overline\alpha(a)} -
1528: X (p_\alpha+a \delta p /2) \sigma_+\big)^2}
1529: {(\delta \zeta_\alpha)^2 + (\delta \zeta_{\overline\alpha(a)})^2}
1530: \eeq
1531: and for each value of $a$ we calculate the optimal value of
1532: $X$, $\overline X(a)$,
1533: by minimizing $\chi^2(a,X)$. The function $\overline X(a)$ is reported in
1534: Fig.~\ref{fig_5}. As the shift of the maximum $\delta$ is between $a=1$ and
1535: $a=2$, we see that the slope $X$ is compatible with one. Moreover, as the
1536: natural error on $p$ is the size of a bin $\delta p$, we assign to the
1537: value $X=1$ a statistical error
1538: $\delta X= 2 ( \overline X(2)-\overline X(1) ) = 0.22$.
1539: Note again that without the translation of $p$ the optimal slope would be
1540: $X \sim 0.85$, incompatible with Eq.~\ref{4}.
1541:
1542: \vskip10pt
1543: \noindent
1544: {\it Discussion} - From the present analysis, we can conclude that:
1545: \begin{enumerate}
1546: \item the translation shifting the maximum of
1547: $\z_\t(p)$ to $p=1$ at the same time
1548: minimizes the difference between $\h_\tau(p)$ and $\h_\tau(-p)+p \sigma_+$,
1549: where $\h_\t$ is our
1550: finite time estimate of $\z_\io$; this proves the consistency of our
1551: theory of finite time corrections;
1552: \item without the translation of $p$ (that corresponds to $a=0$),
1553: the function $\zeta_\tau(p)$ for $\tau \sim 5.0$
1554: {\it do not satisfy the fluctuation relation}, as $\chi(a=0) = 11$
1555: and $\overline X(a=0)=0.85$;
1556: \item the function $\eta_\tau(p)=\zeta_\tau(p-\delta)$
1557: satisfies the fluctuation relation
1558: with $\chi \sim 3$ and an error of about $20\%$ on the slope $X$:
1559: both quantities measure
1560: the accuracy of our analysis.
1561: \end{enumerate}
1562: Thus, the check of the fluctuation relation relies
1563: crucially on the translation
1564: of the function $\zeta_\tau(p)$ that has been discussed
1565: in section~\ref{sec:II}.
1566: By considering larger values of $\tau$ one could avoid this problem
1567: (as $\delta \sim \tau^{-1}$); however,
1568: as one can see from Fig.~\ref{fig_1}, for $\tau > 5.0$
1569: the negative tails of $\zeta_\tau(p)$ are not accessible to our computational
1570: resources. The computation of the finite time corrections
1571: is mandatory if one aims to test
1572: the fluctuation relation at high values of the external driving force.
1573:
1574: \vskip10pt
1575: \noindent
1576: {\it Summary of the data analysis} - To conclude, we summarize the procedure
1577: we follow to analyze the data of a given simulation run:
1578: \begin{enumerate}
1579: \item we determine a value of $\tau$ such that $\zeta_\tau(p)$ appear to
1580: be close to the asymptotic
1581: limit $\zeta_\infty(p)$;
1582: \item we determine the maximum $\widetilde p$ of $\zeta_\tau(p)$
1583: by a sixth-order polynomial fit around $p=1$, in an
1584: interval as big as possible compatibly with the request
1585: that the $\c^2$ from the fit is less than $\sim 10$;
1586: \item we shift the histogram of an integer multiple $a$ of the half
1587: bin size $\delta p/2$ and
1588: compute the function $\chi(a)$ according to Eq.~\ref{chia}.
1589: We determine the value $a^*$
1590: such that the minimum of $\chi(a)$ is
1591: assumed in the interval $[a^*,a^*+1]$:
1592: the consistency of our analysis requires that
1593: $\delta = 1 - \widetilde p$ and $\d_0=(a^*+0.5)\d p/2$
1594: coincide within their errors
1595: (\ie $|\d-\d_0|<\d p$);
1596: \item The value $\chi^* = \min [\chi(a^*),\chi(a^*+1)]$ is an
1597: upper limit for the value of $\chi$ at the minimum.
1598: The number of bins $\min\{M_{a^*},M_{a^*+1}\}$ involved in
1599: this estimate will be called $M^*$;
1600: \item we compute the error $\delta X = 2( \overline X(a^*+1) -
1601: \overline X(a^*))$.
1602: \end{enumerate}
1603: The relevant quantities $\tau$, $\delta$, $\d_0$, $|\d-\d_0|$, $\d p$,
1604: $M^*$, $\chi^*$ and $\delta X$
1605: for model I are
1606: reported in table~\ref{tab:I} for different values of the external force $E$.
1607:
1608: \section{Numerical simulation of model I}
1609:
1610: \begin{figure}[t]
1611: \includegraphics[width=.50\textwidth,angle=0]{FIG6.eps}
1612: \caption{Model I: mobility $\mu$ as a function of the driving force $E$.
1613: The full line is the
1614: equilibrium diffusion coefficient $D$ divided by the temperature.
1615: Deviations from the
1616: linear response are observed around $E=5$.
1617: The error bars are of the order of the
1618: dimension of the symbols. Studying $\m(E)$
1619: for values of $E$ bigger than those shown in the figure,
1620: one can verify that
1621: the mobility increases
1622: up to a value $\m_{max}$, reached in correspondence of
1623: $E\sim 45$. For values of $E$ bigger than $E\sim 45$, the mobility
1624: begins to decrease essentially following the limiting curve
1625: $T J_T/(NE)$, where $J_T=\sqrt{T(d-1/N)}N/T$
1626: is the maximum allowed value of the current (saturation value).}
1627: \label{fig_6}
1628: \end{figure}
1629:
1630: \begin{table}[b]
1631: \begin{tabular}{r|ccccccccc}
1632: \hline
1633: $E$ & $\tau$ & $\sigma_+$ & $\delta$ & $\d_0$ & $|\d-\d_0|$ &
1634: $\d p$ & $M^*$ & $\chi^*$ & $\delta X$ \\
1635: \hline
1636: 2.5 & 5.0 & 0.194 & 0.272 & 0.183 & 0.089 & 0.244 & 43 & 2.2 & 0.24 \\
1637: 5.0 & 5.0 & 0.810 & 0.215 & 0.139 & 0.076 & 0.187 & 20 & 3.5 & 0.22 \\
1638: 7.5 & 4.0 & 1.945 & 0.197 & 0.116 & 0.081 & 0.116 & 18 & 2.8 & 0.18 \\
1639: 10.0 & 2.5 & 4.044 & 0.262 & 0.151 & 0.111 & 0.122 & 17 & 4.4 & 0.20 \\
1640: 12.5 & 2.5 & 7.090 & 0.257 & 0.137 & 0.120 & 0.111 & 8 & 3.5 & 0.28 \\
1641: \hline
1642: \end{tabular}
1643: \caption{Model I: results of the data analysis
1644: for some selected values of $E$.
1645: All the quantities are defined in section~\ref{sec:IV}.
1646: For $E > 12.5$ the negative tails of the distribution
1647: are not accessible to our numerical simulation.}
1648: \label{tab:I}
1649: \end{table}
1650:
1651: We will now discuss systematically the numerical data
1652: obtained from the simulation
1653: of model I (defined in section~\ref{sec:III})
1654: at different values of the driving
1655: force $E$. In Fig.~\ref{fig_6} we report the {\it mobility}
1656: $\mu(E) = T \langle J \rangle_E / (N E)$, {\it i.e.}
1657: the l.h.s. of Eq.~\ref{GKestesa} times $T/N$, as
1658: a function of $E$. The current $J(\ul p,\ul q)$ has
1659: been defined in Eq.~\ref{Jcolor}.
1660: From the Green-Kubo relation, Eq.~\ref{GK},
1661: we have \cite{EM90}
1662: \beq
1663: \lim_{E \rightarrow 0} \mu(E) = \frac{D}{T} \ ,
1664: \eeq
1665: where $D$ is the equilibrium diffusion coefficient,
1666: \beq
1667: D= \lim_{t\rightarrow \infty} \frac{1}{2Nd} \sum_i
1668: \langle |q_i(t)-q_i(0)|^2 \rangle_{E=0}
1669: \eeq
1670: Deviations from the linear response are observed and
1671: $\mu(E) \sim D/T + O(E^2)$ above $E=5$.
1672:
1673: \begin{figure}[t]
1674: \includegraphics[width=.40\textwidth,angle=0]{FIG7.eps}
1675: \caption{Model I: Lyapunov exponents for $E=5$ (top) and for $E=25$ (bottom).
1676: For each panel, the upper and lower dots are the two paired
1677: exponents $\lambda^{(+)}_j$
1678: and $\lambda^{(-)}_j$, and the
1679: middle dot is their average $(\lambda^{(+)}_j+\lambda^{(-)}_j)/2$.
1680: The full line is $\sigma_+/2Nd$, the dashed line is at $\lambda=0$.
1681: }
1682: \label{fig_7}
1683: \end{figure}
1684:
1685: In table~\ref{tab:I} we report the main parameters
1686: that result from the data analysis
1687: (as discussed in the previous section) for some selected values of $E$.
1688: The value $|\d-\d_0|$ is always less than $\d p$, consistently with our
1689: discussion above, except for $E=12.5$ where, however, the relative difference
1690: between the two quantities is small ($\sim 9\%$).
1691: It can be noted that
1692: $\d$ is systematically bigger than $\d_0$.
1693: This could be due to the fact that
1694: the error terms $O((p-1)^2/\t)$ or $o(1/\t)$ that we are discarding
1695: likely produce a systematic shift in $\d$ or in $\d_0$; or that
1696: the velocity of convergence of $\z_\t(p)$ is not the same on the negative or
1697: on the positive side (because numerically is much more difficult to
1698: observe big negative fluctuations of $\s$ than the positive ones --
1699: and the Fluctuation Relation provides a quantitative estimate
1700: of the relative probabilities).
1701: At the moment, because of the level of precision of our
1702: simulations, we are not able to investigate this problem in more
1703: detail, see also Remark (3) in section~\ref{sec:IID}.
1704: On increasing the value of $E$, we are forced to decrease
1705: the value of $\tau$ we use for
1706: the analysis as, for longer $\tau$,
1707: the negative tail of the distribution $\zeta_\tau(p)$
1708: becomes unobservable. This can be seen as the number
1709: $M^*$ of bins used for the computation of
1710: $\chi$ decrease on increasing $E$;
1711: above $E=12.5$ it is impossible to find a value of
1712: $\tau$ such that $\zeta_\tau(p)$
1713: is close to the asymptotic limit and the negative tail
1714: is observable. Thus, the fluctuation relation
1715: cannot be tested above $E=12.5$ with
1716: our computational power. However, we are able to
1717: check the fluctuation relation in
1718: the region $E > 5$ where deviations from the
1719: linear response are observed. Moreover,
1720: the estimated distributions $\zeta_\io(p)$ are very similar to the one reported
1721: in Fig.~\ref{fig_3}: in particular, they are not Gaussian
1722: in the investigated
1723: interval of $p$ (also for $E < 5$, in the linear response regime).
1724:
1725: Finally, in Fig.~\ref{fig_7}
1726: we report the measured Lyapunov exponents of the model for $E=5$
1727: and $E=25$. For this system, the Lyapunov exponents are
1728: known to be paired \cite{SEC98,SEI98,DM96}
1729: like in Hamiltonian systems
1730: and the average of each pair is a constant equal to $\sigma_+/2Nd$.
1731: For $E=5$, each pair is composed of a negative and a positive exponent.
1732: This means that the attractive set is dense in phase space \cite{BGG97,Ga99b}
1733: and the chaotic hypothesis is expected to apply to the system yielding a slope
1734: $X=1$ in the fluctuation relation, as confirmed by our numerical data. The same
1735: happens up to $E \sim 20$.
1736: Above $E=20$, there is a number $D$ of pairs composed by two negative exponents
1737: (for $E=25$ we get $D=4$, see Fig.~\ref{fig_7}).
1738: In this situation, the slope $X$ in the fluctuation relation is expected to be
1739: given by $X = 1 - D/Nd$ \cite{BG97,Ga99b}.
1740: Thus, for $E=25$ one expects $X \sim 0.75$. Unfortunately,
1741: as discussed above, above $E=12.5$ we did not observe negative fluctuations
1742: of the entropy production, and this prediction could not be tested in
1743: our simulation.
1744:
1745:
1746: \section{Numerical simulation of model II}
1747:
1748: \begin{figure}[t]
1749: \includegraphics[width=.50\textwidth,angle=0]{FIG8.eps}
1750: \caption{Mobility as a function of the temperature $T$ and of the
1751: driving force $E$ for model II. The circles correspond to the
1752: equilibrium diffusion coefficient divided by the temperature.
1753: Deviations from the linear response are observed for $E \geq 3$;
1754: they become larger on lowering the temperature,
1755: as $D \rightarrow 0$.
1756: }
1757: \label{fig_8}
1758: \end{figure}
1759:
1760: \begin{figure}[t]
1761: \includegraphics[width=.50\textwidth,angle=0]{FIG9.eps}
1762: \caption{The estimate of the function $\zeta_\io(p)$ (open circles) for
1763: model II with $T=1.1$ and $E=3$. In the same
1764: plot $\zeta_\io(-p) + p\sigma_+$ (filled squares) is reported.
1765: In the inset, the
1766: interval $p\in[-1.5,1.5]$ where the data overlap is magnified. The full
1767: line is the Gaussian approximation, $\zeta_\io(p) =
1768: \frac12\zeta_\io^{(2)}(p-1)^2$.
1769: The data have been obtained from the histogram of $\pi_\tau(p)$ with $\tau=2.5$
1770: (see table~\ref{tab:II}).
1771: }
1772: \label{fig_9}
1773: \end{figure}
1774:
1775: \begin{table}[b]
1776: \begin{tabular}{rr|ccccccccc}
1777: \hline
1778: $T$ & $E$ & $\tau$ & $\sigma_+$ & $\delta$ & $\d_0$ & $|\d-\d_0|$ &
1779: $\d p$ & $M^*$ & $\chi^*$ & $\delta X$ \\
1780: \hline
1781: 0.9 & 1 & 3.0 & 0.209 & 0.453 & 0.334 & 0.119 & 0.223 & 68 & 1.9 & 0.19 \\
1782: 0.9 & 3 & 3.0 & 2.615 & 0.286 & 0.264 & 0.024 & 0.132 & 15 & 1.0 & 0.23 \\
1783: \hline
1784: 1.1 & 1 & 4.0 & 0.233 & 0.231 & 0.126 & 0.105 & 0.126 & 79 & 1.7 & 0.24 \\
1785: 1.1 & 3 & 2.5 & 2.493 & 0.217 & 0.238 & 0.021 & 0.087 & 30 & 1.0 & 0.12 \\
1786: 1.1 & 6 & 1.5 & 13.32 & 0.113 & 0.230 & 0.117 & 0.092 & 7 & 1.1 & 0.21 \\
1787: \hline
1788: 1.5 & 1 & 3.0 & 0.230 & 0.179 & 0.140 & 0.039 & 0.140 & 86 & 0.9 & 0.13 \\
1789: 1.5 & 3 & 2.5 & 2.227 & 0.145 & 0.123 & 0.022 & 0.082 & 33 & 4.7 & 0.18 \\
1790: 1.5 & 6 & 0.5 & 52.14 & 0.074 & 0.130 & 0.056 & 0.052 & 11 & 0.6 & 0.10 \\
1791: \hline
1792: 1.7 & 1 & 3.0 & 0.221 & 0.127 & 0.141 & 0.014 & 0.283 & 49 & 1.0 & 0.26 \\
1793: \hline
1794: 1.9 & 3 & 2.5 & 1.981 & 0.106 & 0.122 & 0.016 & 0.122 & 26 & 0.8 & 0.12 \\
1795: 1.9 & 6 & 0.4 & 43.52 & 0.078 & 0.126 & 0.048 & 0.085 & 14 & 1.7 & 0.11 \\
1796: 1.9 & 10 & 0.2 & 139.0 & 0.079 & 0.135 & 0.056 & 0.039 & 7 & 0.8 & 0.10 \\
1797: \hline
1798: 2.1 & 6 & 0.4 & 40.48 & 0.074 & 0.110 & 0.036 & 0.110 & 11 & 1.0 & 0.15 \\
1799: \hline
1800: \end{tabular}
1801: \caption{Model II: results of the data analysis
1802: for some selected values of $T$ and $E$.
1803: All the quantities are defined in section~\ref{sec:IV}.
1804: }
1805: \label{tab:II}
1806: \end{table}
1807:
1808: Model II differs from model I in the dimension $d=3$, in the larger
1809: number of particles $N=20$, and because it is a binary mixture of
1810: two types of particles.
1811: Binary mixtures are frequently used as models for numerical simulations
1812: of supercooled liquids as they avoid crystallization also
1813: at very low temperature on the "physical" time scales (\ie on the time scales
1814: of numerical experiments); for these systems, at low temperature
1815: deviations from the linear response are observed
1816: also for very low values of the external driving force.
1817:
1818: In Fig.~\ref{fig_8} we report the equilibrium diffusion coefficient
1819: $D$ (divided by the temperature $T$) and the mobility (for different
1820: values of $E$) as functions of the temperature.
1821: Even though the number of particles is very small, on lowering the
1822: temperature the systems approaches the supercooled
1823: state and $D$ becomes very small around $T \sim 0.5$.
1824: Slightly above this temperature, \ie around $T=1$,
1825: strong deviations from the linear response are
1826: observed for $E \geq 3$, where the entropy production $\sigma_+$ is still
1827: close to $0$. Some values of $\sigma_+$ are reported in table~\ref{tab:II};
1828: to compare these values with those obtained for model I one should note
1829: that $\sigma_+$ is an {\it extensive} quantity. Thus, the entropy production
1830: {\it per degree of freedom}, $\sigma_+/2Nd$, is much smaller in model II
1831: than in model I.
1832:
1833: In table~\ref{tab:II} the results of the data analysis outlined in
1834: section~\ref{sec:IV} are reported. For $E \leq 6$ we obtain a very good
1835: agreement of the data with the predictions of the fluctuation relation
1836: and with the theory of finite time
1837: corrections discussed in section~\ref{sec:II}.
1838: For $E=10$ it is very difficult to observe negative fluctuations of $p$
1839: with our computational power; see {\it e.g.} the result of the
1840: analysis for $E=10$ and
1841: $T=1.9$, where only $M^*=7$ bins where available and we were forced to use
1842: $\tau=0.2$, of the order of the mixing time $\tau_0$.
1843: In Fig.~\ref{fig_9} we report the estimated function $\zeta_\io(p)$ obtained
1844: for $T=1.1$ and $E=3$ from the data with $\tau=2.5$. Strong deviations from
1845: the Gaussian behavior are observed in the accessible range of $p$ (see the
1846: inset of Fig.~\ref{fig_9}). A similar behavior
1847: of $\z_\io(p)$ is observed in correspondence of all the values
1848: of $E$ and $T$ we investigated (those listed in Table II):
1849: in particular in all these cases highly non Gaussian behaviors
1850: are observed in the accessible range of $p$.
1851:
1852: The Lyapunov spectrum for this system is very similar to the one reported
1853: in the upper panel of Fig~\ref{fig_7}. Pairs of two negative exponents were
1854: observed only for $E=10$ at $T \leq 1.3$, where, as in the case of model I,
1855: $\sigma_+$ is too large to allow for a
1856: verification of the modified fluctuation relation expected in this case,
1857: see the discussion at the end of section V.
1858:
1859: \section{Conclusions}
1860:
1861: We tested the fluctuation relation, in our opinion quite successfully,
1862: in a numerical simulation of
1863: two models of interacting particles subjected to an external nonconservative
1864: force and to a reversible mechanical thermostat.
1865: Our data satisfy the fluctuation relation with a $\chi \leq 3$
1866: and an accuracy of the order of $20 \%$ also for very large values of the
1867: driving force, where strong deviations from the linear response are observed,
1868: and where the large deviation function is strongly non-Gaussian.
1869: The comparison of our numerical data with the predictions of the
1870: fluctuation relation is done by taking into account the
1871: (lowest order) finite time
1872: corrections to the distribution function for the fluctuations of
1873: the phase space contraction rate. This is crucial: if we did not
1874: take into account such corrections the fluctuation relation would
1875: be violated within the precision of our experiment.
1876:
1877: In order to compute the finite time corrections,
1878: we proposed an algorithm which allows to reconstruct the asymptotic
1879: distribution function from measurable quantities at finite time, within
1880: a given precision. Our theory of the corrections relies on the
1881: symbolic representation of the chaotic dynamics, therefore it is
1882: applicable if one accepts the Chaotic Hypothesis.
1883:
1884: Our interpretation of the numerical results is that the
1885: {\it chaotic hypothesis} can be applied to these
1886: systems, also very far from equilibrium, and in particular
1887: the fluctuation relation is
1888: satisfied even in regions where its predictions measurably differ
1889: from those of linear response theory.
1890:
1891: Our theory of finite time corrections for the analysis of our
1892: numerical data could in principle be of interest for real experimental
1893: settings where non Gaussian fluctuations for the entropy production
1894: rate are observed, see \cite{FM04,CL98}.
1895:
1896: However it should be stressed that in a real experiment there are some
1897: technical differences with respect to our numerical simulation which
1898: could in some cases make
1899: inapplicable
1900: our analysis, namely:
1901: \\
1902: {\it (i)} usually the noise in the large deviation function for the
1903: entropy production rate in a real experiment is much bigger than in a
1904: numerical experiment, and it is likely that the translation in
1905: Eq.~\ref{etatau} computed as the ratio $\z^{(3)}/(\z^{(2)})^2$ is not
1906: measurable within an error of some percent;
1907: \\
1908: {\it (ii)} usually in a real experiment the accessible time scales
1909: are naturally much bigger than the microscopic ones so that, if the
1910: negative fluctuations of the entropy production rate are observable at all,
1911: one is automatically in the asymptotic regime, where the finite time
1912: corrections should be negligible;
1913: \\
1914: {\it (iii)} a usual problem in a realistic setting is that there is no
1915: clear connection between the ``natural'' thermodynamic entropy
1916: production rate $\dot s=W/T$ ($W$ is the work of the dissipative
1917: external forces and $T$ is the temperature) and the microscopic phase
1918: space contraction rate, for which a slope $X=1$ in the fluctuation
1919: relation $\z(p)-\z(-p)=X\sigma_+p$ is expected;
1920: so, often one measures an $X\neq 1$ and correspondingly one {\it
1921: defines} an effective temperature $T_{eff}=T/X$ giving a natural
1922: connection between the effective thermodynamic entropy production rate
1923: $\dot s_{eff}=W/T_{eff}$ and the phase space contraction rate, see
1924: \cite{CL98,FM04,Ga04};
1925: in such a situation (where an adjustable parameter $X$ appears) it makes
1926: no sense to apply our analysis, which is sensible only if one wants to compare
1927: the experimental data with a sharp prediction about the slope $X$ in the
1928: fluctuation relation.
1929:
1930: A big open problem we are left with is trying to understand
1931: how the fluctuation relation is modified for values
1932: of the driving force so high that the attractive set is no longer dense in phase
1933: space. It is expected, \cite{BG97},
1934: that in such a case
1935: $\z_\io(p)-\z_\io(-p)$ is still linear, but the slope is $X\s_+$, with
1936: $X$ given by the ratio of the dimension of the attractive set and
1937: of that of the whole phase space. An estimate of such quantity can be given
1938: via the number of negative pairs of exponents in the Lyapunov spectrum
1939: \cite{BG97,Ga99b}.
1940: Unfortunately negative pairs begin to appear in the Lyapunov spectrum
1941: only for values of the external force so high that no negative fluctuations
1942: are observable anymore.
1943: We hope that future work will address this point.
1944:
1945: \appendix
1946: \section{A Limit Theorem}
1947: \label{app:A}
1948:
1949: In this section we prove Eq.~\ref{24}--\ref{23}.
1950: We reproduce in detail the proof
1951: in the case $p$ is the average
1952: of independently distributed discrete variables $\s_i^\e$, assuming
1953: values in $\e\ZZZ$, for some small mesh parameter $\e$; then we discuss
1954: how this can be applied and adapted to the situation
1955: considered in section IIC and subsequent sections.
1956:
1957: Let us introduce some definitions. Let $\s_i$, $i\in\NNN$,
1958: be independent continuous random variables with identical
1959: distributions $\p(d\s_i)$ with positive variance $\d\s^2>0$, supported
1960: on the finite interval $[s_-,s_+]$. Let us assume that $\p(d\s_i)$
1961: gives positive probability to any finite interval contained in $[s_-,s_+]$.
1962: Let $\p_\l(d\s)$ be the weighted distribution
1963: $\p_\l(d\s)=e^{-\l\s}\p(d\s)/\int e^{-\l\s}\p(d\s)$ and let us define
1964: $z_\io(\l)=-\log\int e^{-\l\s}\p(d\s)$ and $\s_+=z_\io'(0)$.
1965: Note that the assumption
1966: that $\p(d\s_i)$ gives positive probability to an interval of $\s$
1967: in $[s_-,s_+]$ implies that for any finite
1968: $\l$ also $\p_\l(d\s)$ has positive variance $-z_\io''(\l)>0$.
1969:
1970: Also, given $\e>0$ (with the property that $s_+-s_-=N_\e\e$ for some integer
1971: $N_\e$),
1972: let us consider the discretization of
1973: $\s_i$ on scale $\e$, call it $\s_i^\e$: $\s_i^\e$ will be a discrete variable
1974: assuming the values
1975: $s_k^\e\defin s_-+(k-\frac{1}{2})\e$, $k=1,\ldots,N_{\e}$,
1976: with probabilities $\p^\e(s_k^\e)=
1977: \Prob(\s_i^\e=s_k^\e)=\int_{s_k^\e\pm\frac{\e}{2}}
1978: \p(d\s)$. The assumption that $\p(d\s_i)$
1979: gives positive probability to any finite interval contained
1980: in $[s_-,s_+]$ implies that $\p^\e(s_k^\e)>0$ for any $\e$ and $k$.
1981: Let also $z_\e(\l)=-\log\sum_{k=1}^{N_\e}e^{-\l s_k^\e}\p^\e
1982: (s_k^\e)$ and $\p^\e_\l(s_k^\e)=\p^\e(s_k^\e)e^{-\l s_k^\e+z_\e(\l)}$.
1983: Note that, since $\p^\e(s_k^\e)>0$ for any $k$, for any finite $\l$ one has
1984: $-z_\e''(\l)>0$.
1985:
1986: If $p_\t^\e=\frac{1}{\t\s_+}\sum_{i=1}^\t\s_i^\e$ and $\P_\t(\e;I)$ is the
1987: probability
1988: that $p_\t^\e$ belongs
1989: to the finite interval $I$, the following theorem holds.\\
1990: \\
1991: {\bf Theorem:}
1992: {\it Given a finite interval $I\subset (s_-,s_+)$, let $\s_i^\e$, $\p^\e$ and
1993: $\P_\t(\e;I)$ be defined as above. Then, for a sufficiently small $\e>0$,
1994: there exists an analytic "rate function" $\widetilde\z_\t(p)$ such that
1995: %
1996: \beq\label{A1.1}
1997: \lim_{\t\to\io}\frac
1998: {\P_\t(\e;I)}
1999: {\int_{I} dp e^{\t\wt\z_\t(p)}}
2000: =1\;.\eeq
2001: %
2002: $\widetilde\z_\t(p)$ is defined by:
2003: %
2004: \beq\label{A1.2}\begin{split}
2005: &\wt\z_\t(p)+\frac{1}{\t}\log\Big[\frac{\sinh[\e\l_p^\e/(2\s_+)]}
2006: {\e\l_p^\e/(2\s_+)}
2007: \Big]=\z_\t^\e(p)\\
2008: &\z_\t^\e(p)=-z_\e(\l_p^\e)+\l_p^\e p\s_+-\frac{1}{2\t}
2009: \log[\frac{2\p}{\t}\Big(-\frac{z_\e''(\l_p^\e)}{\s_+^2}\Big)]\end{split}
2010: \eeq
2011: %
2012: and $\l_p^\e$ is the inverse of $p(\l)=z_\e'(\l)/\s_+$.
2013: The function $\z_\t^\e(p)$ has the
2014: following property: if $\D\subset I$ is an interval of
2015: size $\frac{\e}{\t\s_+}$ around
2016: a point $p_\D$, then:
2017: %
2018: \beq\label{A1.1a}
2019: \lim_{\t\to\io}\frac
2020: {\P_\t(\e;\D)}{|\D|e^{\t\z_\t^\e(p_\D)}}=1\eeq
2021: %
2022: }
2023: \\
2024: {\bf Proof}
2025: %Given $I$, let $J$ be the image of $I$ through $\l=\l_p$ (that is well
2026: %defined if $\e$ is chosen small enough).
2027: Let us introduce the
2028: auxiliary variable $q={1\over\t\s_+}\sum_{i=1}^\t\h_i$,
2029: where $\h_i$ are i.i.d. discrete random variables, with distribution
2030: $\p_\l^\e(s_k^\e)$. Let us call $\P_\t^\l(\e;q_0)$ the probability that
2031: $q$ assumes the value $q_0\in I$, with $q_0\s_+=s_k^\e/\t$ for some $k\in\NNN$,
2032: and note that
2033: $\P_\t^0(\e;q_0)$ is identical to the probability that $p_\t=q_0$.
2034: By definition $\P_\t^\l(q_0)$ and $\P_\t^0(q_0)$ are related by:
2035: %
2036: \beq\label{A1.5}
2037: \P_\t^\l(\e;q_0)={e^{-\l q_0\s_+\t}\P_\t^0(\e;q_0)\over
2038: \big[\sum_k e^{-\l s_k^\e} \p^\e(s_k^\e)\big]^\t}
2039: \eeq
2040: %
2041: Now, a local form of central limit theorem
2042: (Gnedenko's theorem, see pag. 211 of \cite{Fi63}) tells us that,
2043: if $q$ is localized near its mean value, that is if
2044: $|q\s_+-z_\e'(\l)|\le \frac{M\e}{\t}$
2045: for some finite $M$, then $\P_\t^\l(\e;q_0)$ is asymptotically equivalent
2046: to the Gaussian with mean $z_\e'(\l)$ and variance
2047: $-z_\e''(\l)$, in the sense that
2048: %
2049: \beq\label{A1.6}\P_\t^\l(q_0)=\frac{\e}{\sqrt{2\p\t(-z_\e''(\l))}}e^{-
2050: \frac{(q_0\s_+-z'_\e(\l))^2}{2(-z_\e''(\l))}\t}(1+o(1))\;,\eeq
2051: %
2052: for any $q_0$ s.t. $|q\s_+-z_\e'(\l)|\le \frac{M\e}{\t}$
2053: \footnote{\label{central} Note that Gnedenko's Theorem is {\it different}
2054: from the usual central limit theorem, stating instead that for
2055: $|q\s_+-z_\e'(\l)|\le \frac{C}{\sqrt\t}$ ($C$ big) the sums of
2056: $\P_\t^\l(\e;q)$ over intervals of amplitude $\frac{1}{\sqrt\t}$ contained in
2057: $|q\s_+-z_\e'(\l)|\le \frac{C}{\sqrt\t}$
2058: are asymptotically equal to the integrals of the Gaussian
2059: over the same intervals.
2060: That is, usual central limit theorem gives informations on the distribution
2061: in a bigger interval around the maximum, but on a rougher scale.}.
2062:
2063: So, given $\l_{q_0}^\e$ s.t. $z_\e'(\l_{q_0}^\e)=q_0\s_+$ (such $\l_{q_0}^\e$
2064: exists, is unique
2065: and is an analytic function of $q_0$, by the remark that $-z_\e''(\l)>0$ for
2066: any finite $\l$ and
2067: $z_\e(\l)$ is an analytic function of $\l$), using
2068: Eq.~\ref{A1.6} we see that
2069: Eq.~\ref{A1.5} can be restated as:
2070: %
2071: \beq\label{A1.7}\P_\t^0(\e;q_0)=\frac{\e}{\sqrt{2\p\t(-z_\e''(\l_{q_0}^\e))}}
2072: e^{\l_{q_0}^\e q_0\s_+\t-z_\e(\l_{q_0}^\e)}(1+o(1))\eeq
2073: %
2074: Now, by the definition of $\z_\t^\e(p)$ in Eq.~\ref{A1.2}, we see that
2075: the r.h.s. of the last equation is equal to
2076: $\frac{\e}{\t\s_+}e^{\t\z_\t^\e(q_0)}(1+o(1))$.
2077: Finally, the statement of the Theorem follows by the remark that
2078: %
2079: \beq\label{A1.8}
2080: \frac{\e}{\t\s_+}e^{\t\z_\t^\e(p_0)}=\int_{p_0-\frac{\e}{2\t\s_+}}^{p_0+\frac{\e}
2081: {2\t\s_+}}dp e^{\t\tilde\z_\t(p)}\Big(1+o(1)\Big)\;.\eeq
2082: %
2083: In fact the integral in the r.h.s. of the last equation is given by
2084: %
2085: \beq \label{A1.9}\begin{split}
2086: &e^{\t\tilde\z_\t(p_0)} \int_{p_0-\frac{\e}{2\t\s_+}}^{p_0+\frac{\e}
2087: {2\t\s_+}}dp e^{\t\tilde\z_\t'(p_0)(p-p_0)}\Big(1+O(\frac{\wt\z_\t''(p_0)\e^2}{\t})
2088: \Big)=\\
2089: &=e^{\t\tilde\z_\t(p_0)} \frac{2\sinh[\tilde\z_\t'(p_0)\e/(2\s_+)]}
2090: {\t\tilde\z_\t'(p_0)}
2091: \Big(1+O(\frac{\wt\z_\t''(p_0)\e^2}{\t})\Big)\end{split}\eeq
2092: %
2093: and in the last expression one has to note that $\wt\z_\t'(p_0)=[\z_\t^\e]'(p_0)+
2094: O(\frac{1}{\t})=\l_{p_0}^\e+O(\frac{1}{\t}).\ $ \qed
2095: \\
2096: \\
2097: A first Remark to be done about the Theorem above is that, in order
2098: to define a ``universal'' rate function in terms of quantities
2099: depending only on
2100: $z_\io(\l)$ (instead of quantities depending
2101: on the ``non universal'' function $z_\e(\l)$,
2102: which explicitly depends on the discretization step $\e$), it would
2103: desirable to perform (in a sense to be precised)
2104: the continuum limit $\e\to 0$. To this regard, we can note that
2105: the only point where in the proof above we really used the fact that
2106: $\e$ is a constant (\ie is independent of $\t$) was in using Gnedenko's
2107: Theorem, see \cite{Fi63}. However, by a critical analysis of the proof
2108: of Gnedenko's Theorem, one can realize that it is even possible to
2109: let $\e=\e_\t$ go to $0$ with $\t$;
2110: the velocity with which $\e_\t$ is allowed to go to $0$ depends on the
2111: details of the distribution $\p(d\s)$.
2112: So we can even study the probability distribution of $p_\t$ on a scale
2113: $\sim \e_\t/\t$: if we introduce bins $\D_\t$ of size $O(\e_\t/\t)$
2114: and we define $\P_\t(\D_\t)$ to be the probability that $p_\t={1\over \t\s_+}
2115: \sum_i\s_i$ belongs to the bin $\D_\t$ centered in $p_0$, we can
2116: repeat the proof above to conclude that
2117: %
2118: \beq\label{A1.9a}
2119: \lim_{\t\to\io}\frac
2120: {\P_\t(\D_\t)}{|\D_\t|e^{\t\z_\t(p_0)}}=1\eeq
2121: %
2122: where $\z_\t$ satisfies the equation:
2123: %
2124: \beq\z_\t(p)=-z_\io(\l_p)+\l_p p\s_+-\frac{1}{2\t}
2125: \log[\frac{2\p}{\t}\Big(-\frac{z_\io''(\l_p)}{\s_+^2}\Big)]\eeq
2126: %
2127: and $\l_p$ is the inverse of $p(\l)=z_\io'(\l)/\s_+$. \\
2128:
2129: Another point to be discussed is that in the Theorem above we assumed
2130: the $\s_i$ to be independent.
2131: This is not the case for the variables $\s(S^i\cdot)$ of sec.~\ref{sec:II}.
2132: However, if, as discussed in Remark (3) of sec.~\ref{sec:IID},
2133: we choose the time unit to be of the order of the mixing time, the variables
2134: $\s(S^i\cdot)$ have (by construction) a decorrelation time equal to $1$, and the
2135: analysis of previous theorem can be repeated step by step
2136: in order to construct the probability distribution of
2137: $p=\frac{1}{\t\s_+}\sum_{i}\s(S^i\cdot)$. The only differences
2138: are that: (1) $\t z_\io(\l)$ should be replaced by $\t z_\t(\l)
2139: =-\log\int
2140: e^{-\l p\s_+\t}\P_\t(dp)$ throughout the discussion; (2) instead of Gnedenko's
2141: theorem one has to apply a generalization of Gnedenko's to short ranged
2142: Gibbs processes, to be proven via standard cluster expansion
2143: techniques (see for instance \cite{Ga72} for a proof of a generalization
2144: of Gnedenko's theorem to a short ranged Gibbs process
2145: in the context of non critical fluctuations
2146: of the phase separation line in the 2D Ising model). \\
2147:
2148: The conclusion is that, if the bins $\D$
2149: in sec.~\ref{sec:IIC} are chosen of size $\e_\t/\t$, the probability of the bin $\D$
2150: centered in $p_\D$ is asymptotically given by $\p(p\in\D)\simeq e^{\t\z_\t(p_\D)}$
2151: (in the sense of Eq.~\ref{24}) and $\z_\t(p_\D)$ can be interpolated by
2152: an analytic function of $p$ that in fact satisfies Eq.~\ref{23}.
2153:
2154: \acknowledgments
2155:
2156: We would like to thank Prof.~Joel~L.~Lebowitz for his interest
2157: on this work and for a valuable discussion. Some of this work was
2158: completed while two of us (A.G. and G.G.) were visiting Rutgers University
2159: under invitation of Prof.~Lebowitz. A.G. was partially supported by
2160: the NSF grant 4-23421 DMR 01-279-26.
2161:
2162: We thank F.~Bonetto for a useful discussion and his suggestions about
2163: the implementation of the numerical algorithm.
2164:
2165: F.Z. wish to thank G.~Ruocco and L.~Angelani for many useful
2166: discussions.
2167:
2168: The computations have been performed on the FDT cluster of the INFM-SOFT center
2169: in Rome. We thank S.~Erriu for technical assistance in operating the cluster.
2170: The sources of the program we used for the simulations
2171: can be downloaded from {\it http://glass.phys.uniroma1.it/zamponi}.
2172:
2173: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2174: % REFERENCES
2175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2176:
2177: \begin{thebibliography}{99}
2178:
2179: \bibitem{Ga99} G.~Gallavotti, {\it Statistical mechanics. A short treatise}
2180: (Springer Verlag, Berlin, 1999).
2181:
2182: \bibitem{GC95a} G.~Gallavotti and E.G.D.~Cohen,
2183: {\it Dynamical ensembles in nonequilibrium statistical mechanics},
2184: Phys.~Rev.~Lett. {\bf 74}, 2694--2697 (1995).
2185:
2186: \bibitem{GC95b} G.~Gallavotti, E.~G.~D.~Cohen, {\it Dynamical
2187: ensembles in stationary states}, J.~Stat.~Phys. {\bf 80}, 931--970 (1995).
2188:
2189: \bibitem{Ga95c} G.~Gallavotti, {\it Reversible Anosov maps and large
2190: deviations},
2191: Mathematical Physics Electronic Journal, MPEJ, {\bf 1}, 1--12 (1995).
2192:
2193: \bibitem{Ru99} D.~Ruelle,
2194: {\it Smooth dynamics and new theoretical ideas
2195: in non-equilibrium statistical mechanics},
2196: J.~Stat.~Phys. {\bf 95}, 393--468 (1999).
2197:
2198: \bibitem{ECM93} D.~J.~Evans, E.~G.~D.~Cohen, and G.~P.~Morriss,
2199: {\it Probability of second law violations in shearing steady states},
2200: Phys.~Rev.~Lett. {\bf 71}, 2401--2404 (1993).
2201:
2202: \bibitem{Ga04} G.~Gallavotti, {\it Entropy production in
2203: nonequilibrium stationary states: a point of view},
2204: Chaos, {\bf 14}, 680--690, (2004).
2205:
2206: \bibitem{Ru04} D.~Ruelle, {\it Conversations on nonequilibriun physics with an
2207: extraterrestrial}, Physics Today, May, 48--53 (2004).
2208:
2209: \bibitem{EM90} D.J.~Evans and G.P.~Morris, {\it Statistical Mechanics of
2210: Nonequilibrium Liquids} (Academic Press, London, 1990).
2211:
2212: \bibitem{BGG97} F.~Bonetto, G.~Gallavotti, and P.~L.~Garrido,
2213: {\it Chaotic principle: an experimental test},
2214: Physica D {\bf 105}, 226--252 (1997).
2215:
2216: \bibitem{BCL98} F.~Bonetto, N.~I.~Chernov, J.~L.~Lebowitz,
2217: {\it (Global and local) fluctuations of phase-space contraction
2218: in deterministic stationary non-equilibrium},
2219: Chaos {\bf 8}, 823--833 (1998).
2220:
2221: \bibitem{BPV98} L.~Biferale, D.~Pierotti, and A.~Vulpiani,
2222: {\it Time-reversible dynamical systems for turbulence},
2223: J.~Phys.~A:~Math.~Gen. {\bf 31}, 21--32 (1998).
2224:
2225: \bibitem{GP99} G.~Gallavotti, F.~Perroni,
2226: {\it An experimental test of the local fluctuation theorem in chains of
2227: weakly interacting Anosov systems},
2228: {\it preprint} chao-dyn/9909007.
2229:
2230: \bibitem{GRS04} G.~Gallavotti, L.~Rondoni, and E.~Segre,
2231: {\it Lyapunov spectra and nonequilibrium ensembles equivalence in 2D
2232: fluid mechanics},
2233: Physica D {\bf 187}, 338--357 (2004).
2234:
2235: \bibitem{ZRA04} F.~Zamponi, G.~Ruocco, L.~Angelani,
2236: {\it Fluctuations of entropy production in the isokinetic ensemble},
2237: J.~Stat.~Phys. {\bf 115}, 1655--1668 (2004).
2238:
2239: \bibitem{GR97} G.~Gallavotti, D.~Ruelle, {\it SRB states and nonequilibrium
2240: statistical
2241: mechanics close to equilibrium}, Comm.~Math.~Phys. {\bf 190}, 279--285 (1997).
2242:
2243: \bibitem{Ga96} G.~Gallavotti,
2244: {\it Extension of Onsager's reciprocity to large fields and
2245: the chaotic hypothesis},
2246: Phys.~Rev.~Lett. {\bf 77}, 4334--4337 (1996);
2247: G.~Gallavotti,
2248: {\it Chaotic hypothesis: Onsager reciprocity and
2249: fluctuation-dissipation theorem},
2250: J.~Stat.~Phys. {\bf 84}, 899 (1996).
2251:
2252: \bibitem{Si68} Y.~G.~Sinai, {\it Markov partitions and $C$-diffeomorphisms},
2253: Functional Analysis and Applications, {\bf 2}, n.1, 64--89 (1968);
2254: {\it Construction of Markov partitions}, Functional analysis and
2255: Applications, {\bf 2}, n.2, 70--80, (1968).
2256:
2257: \bibitem{Si77} Y.~G.~Sinai, {\it Lectures in ergodic theory}, Lecture notes
2258: in Mathematics, Prin\-ce\-ton U. Press, Princeton, 1977.
2259:
2260: \bibitem{Ga02} G.~Gallavotti,
2261: {\it Fluid mechanics. Foundations},
2262: Springer-Verlag, Berlin, 2002.
2263:
2264: \bibitem{GBG04} G.~Gallavotti, F.~Bonetto, G.~Gentile,
2265: {\it Aspects of the ergodic, qualitative and statistical properties
2266: of motion}, Springer Verlag, Berlin, 2004.
2267:
2268: \bibitem{DGM} S.~de~Groot, P.~Mazur: {\it Non equilibrium
2269: thermodynamics}, Dover, 1984, (reprinted).
2270:
2271: \bibitem{Ru00} D.~Ruelle, {\it A remark on the equivalence of isokinetic
2272: and isoenergetic thermostats in the thermodynamic limit},
2273: J.~Stat.~Phys. {\bf 100}, 757--763 (2000).
2274:
2275: \bibitem{AT87} M.~P.~Allen and D.~J.~Tildesley, {\it Computer simulation of
2276: liquids} (Oxford Science Publications, 1987).
2277:
2278: \bibitem{SEC98} S.~S.~Sarman, D.~J.~Evans, P.~T.~Cumming,
2279: {\it Recent developments in non-Newtonian molecular dynamics},
2280: Phys.~Rep. {\bf 305}, 1--92 (1998).
2281:
2282: \bibitem{KA94} W.~Kob and C.~H.~Andersen,
2283: {\it Scaling behavior in the $\beta$-relaxation regime of a
2284: supercooled Lennard-Jones mixture},
2285: Phys.~Rev.~Lett. {\bf 73}, 1376--1379 (1994).
2286:
2287: \bibitem{DmSC04} C.~De Michele, F.~Sciortino, A.~Coniglio,
2288: {\it Scaling in soft spheres: fragility invariance on the
2289: repulsive potential softness},
2290: J.~Phys.:~Condens.~Matter {\bf 16}, L489--L494 (2004).
2291:
2292: \bibitem{BGS76} G.~Benettin, L.~Galgani, J.-M.~Strelcyn,
2293: {\it Kolmogorov entropy and numerical experiments},
2294: Phys.~Rev.~A {\bf 14}, 2338--2345 (1976).
2295:
2296: \bibitem{BG97} F.~Bonetto, G.~Gallavotti,
2297: {\it Reversibility, coarse graining and the chaoticity principle},
2298: Communications in Mathematical Physics, {\bf 189}, 263--276, (1997).
2299:
2300: \bibitem{Ga99b} G.~Gallavotti,
2301: {\it New methods in nonequilibrium gases and fluids},
2302: Open Systems and Information Dynamics {\bf 6}, 101--136 (1999).
2303:
2304: \bibitem{SEI98} D.~J.~Searles, D.~J.~Evans, D.~J.~Isbister,
2305: {\it The conjugate-pairing rule for non-Hamiltonian system}, Chaos {\bf 8},
2306: 337--349 (1998).
2307:
2308: \bibitem{DM96} C.~P.~Dettmann and G.~P.~Morriss,
2309: {\it Proof of Lyapunov exponent pairing for systems at
2310: constant kinetic energy},
2311: Phys.~Rev.~E {\bf 53}, R5545--R5548 (1996).
2312:
2313: \bibitem{Fi63} M.~Fisz, {\it Probability theory and mathematical
2314: statistics}, Wiley, New York, 1963.
2315:
2316: \bibitem{Ga72} G.~Gallavotti, {\it Phase separation line in the two--dimensional
2317: Ising model}, Comm. Math. Phys. {\bf 27}, 103--136 (1972).
2318:
2319: \bibitem{CL98} S.~Ciliberto and C.~Laroche,
2320: {\it An experimental test of the Gallavotti-Cohen fluctuation theorem},
2321: J.~Phys.~IV {\bf 8}, Pr6-215 (1998).
2322:
2323: \bibitem{FM04} K.~Feitosa and N.~Menon,
2324: {\it Fluidized Granular Medium as an Instance of the Fluctuation Theorem},
2325: Phys.~Rev.~Lett. {\bf 92}, 164301 (2004).
2326:
2327:
2328: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2329:
2330: \end{thebibliography}
2331:
2332: \end{document}
2333:
2334: