1: %Latex arXiv: 0711.2755v1 (cond-mat.stat-mech)
2: \documentclass[11pt]{article}
3: \title{\bf Heat and Fluctuations from\\ Order to Chaos}
4: \author{\small\textsc{Giovanni Gallavotti}%
5: \thanks{%IUPAP23 Conference, Genova, 2007;
6: Review}\\
7: \small Dipartimento di Fisica and INFN\\
8: \small Universit\`a di Roma
9: {\it La Sapienza}\\
10: \small P.~A.~Moro 2, 00185, Roma, Italy\\
11: \small \texttt{giovanni.gallavotti@roma1.infn.it}
12: }
13: \date{\today
14: }
15: %\date{ November 2007}
16: \newcount\biblio
17: \biblio=1 %0: uses bibtex; 1: uses Ge.bbl
18: \font\ottorm=cmr8\font\ninerm=cmr9%
19: \font\cs=cmcsc10
20: \font\sc=cmcsc10
21: \def\st{\scriptstyle}%
22: \def\dt{\displaystyle}%
23: %
24: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
25: %%%%%%%%%%%%% Greek and Latin/ Footnote Fonts Bold/Small %%%%%%%%%%%%%%%%%%
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27: %
28: % pag.430 TeX manual
29: \font\diecibf=cmbx10%
30: \font\tenmib=cmmib10 \font\eightmib=cmmib8
31: \font\sevenmib=cmmib7\font\fivemib=cmmib5
32: \font\ottoit=cmti8\font\fiveit=cmti5\font\sixit=cmti6%%
33: \font\fivei=cmmi5\font\sixi=cmmi6\font\ottoi=cmmi8%
34: \font\ottorm=cmr8\font\fiverm=cmr5\font\sixrm=cmr6%
35: \font\ottosy=cmsy8\font\sixsy=cmsy6\font\fivesy=cmsy5%%
36: \font\ottobf=cmbx8\font\sixbf=cmbx6\font\fivebf=cmbx5%
37: \font\ottott=cmtt8%
38: \font\ottocss=cmcsc8%
39: \font\ottosl=cmsl8%
40: \def\ottopunti{\def\rm{\fam0\ottorm}\def\it{\fam6\ottoit}\def\em{\fam6\ottoit}%
41: \def\bf{\fam7\ottobf}%
42: \textfont1=\ottoi\scriptfont1=\sixi\scriptscriptfont1=\fivei%
43: \textfont2=\ottosy\scriptfont2=\sixsy\scriptscriptfont2=\fivesy%
44: %\textfont3=\tenex\scriptfont3=\tenex\scriptscriptfont3=\tenex%
45: \textfont4=\ottocss\scriptfont4=\sc\scriptscriptfont4=\sc%
46: \textfont5=\eightmib\scriptfont5=\sevenmib\scriptscriptfont5=\fivemib%
47: \textfont6=\ottoit\scriptfont6=\sixit\scriptscriptfont6=\fiveit%
48: \textfont7=\ottobf\scriptfont7=\sixbf\scriptscriptfont7=\fivebf%
49: \setbox\strutbox=\hbox{\vrule height7pt depth2pt width0pt}%
50: \normalbaselineskip=9pt\rm}
51: \let\nota=\ottopunti%
52: \textfont5=\tenmib\scriptfont5=\sevenmib\scriptscriptfont5=\fivemib
53:
54: \mathchardef\Ba = "050B %alpha
55: \mathchardef\Bb = "050C %beta
56: \mathchardef\Bg = "050D %gamma
57: \mathchardef\Bd = "050E %delta
58: \mathchardef\Be = "0522 %varepsilon
59: \mathchardef\Bee = "050F %epsilon
60: \mathchardef\Bz = "0510 %zeta
61: \mathchardef\Bh = "0511 %eta
62: \mathchardef\Bthh = "0512 %theta
63: \mathchardef\Bth = "0523 %vartheta
64: \mathchardef\Bi = "0513 %iota
65: \mathchardef\Bk = "0514 %kappa
66: \mathchardef\Bl = "0515 %lambda
67: \mathchardef\Bm = "0516 %mu
68: \mathchardef\Bn = "0517 %nu
69: \mathchardef\Bx = "0518 %xi
70: \mathchardef\Bom = "0530 %omega
71: \mathchardef\Bp = "0519 %pi
72: \mathchardef\Br = "0525 %rho
73: \mathchardef\Bro = "051A %varrho
74: \mathchardef\Bs = "051B %sigma
75: \mathchardef\Bsi = "0526 %varsigma
76: \mathchardef\Bt = "051C %tau
77: \mathchardef\Bu = "051D %upsilon
78: \mathchardef\Bf = "0527 %phi
79: \mathchardef\Bff = "051E %varphi
80: \mathchardef\Bch = "051F %chi
81: \mathchardef\Bps = "0520 %psi
82: \mathchardef\Bo = "0521 %omega
83: \mathchardef\Bome = "0524 %varomega
84: \mathchardef\BG = "0500 %Gamma
85: \mathchardef\BD = "0501 %Delta
86: \mathchardef\BTh = "0502 %Theta
87: \mathchardef\BL = "0503 %Lambda
88: \mathchardef\BX = "0504 %Xi
89: \mathchardef\BP = "0505 %Pi
90: \mathchardef\BS = "0506 %Sigma
91: \mathchardef\BU = "0507 %Upsilon
92: \mathchardef\BF = "0508 %Phi
93: \mathchardef\BPs = "0509 %Psi
94: \mathchardef\BO = "050A %Omega
95: \mathchardef\BDpr = "0540 %{\bf\partial}
96: \mathchardef\Bstl = "053F %{\bf*}
97: \def\BK{\bf K}
98:
99: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
100: %%%%%%%%%%%%%%% INSERIMENTO FIGURE ( se si usa DVIPS ) %%%%%%%%%%%%%%
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: \newdimen\xshift \newdimen\xwidth \newdimen\yshift \newdimen\ywidth
103:
104: \def\ins#1#2#3{\vbox to0pt{\kern-#2pt\hbox{\kern#1pt#3}\vss}\nointerlineskip}
105:
106: \def\eqfig#1#2#3#4#5{%
107: \par\xwidth=#1pt\xshift=\hsize\advance\xshift%
108: by-\xwidth\divide\xshift by 2%
109: \yshift=#2pt\divide\yshift by 2%
110: %\line
111: {\hglue\xshift \vbox to #2pt{\vfil%
112: #3\special{psfile=#4}%
113: }\hfill\raise\yshift\hbox{#5}}}%
114:
115: %%%%%%%%%%%%%%%%%%%%%%%%%%
116:
117:
118: %%%%%%%%%%%%%%%%%%%%%% SYMBOLS %%%%%%%%%%%%%%%%%%%%%%%
119: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
120: \let\a=\alpha \let\b=\beta \let\g=\gamma \let\d=\delta \let\e=\varepsilon
121: \let\z=\zeta \let\h=\eta \let\th=\theta \let\k=\kappa \let\l=\lambda
122: \let\m=\mu \let\n=\nu \let\x=\xi \let\p=\pi \let\r=\rho
123: \let\s=\sigma \let\t=\tau \let\f=\varphi \let\ph=\varphi\let\c=\chi
124: \let\ch=\chi \let\ps=\psi \let\y=\upsilon \let\o=\omega\let\si=\varsigma
125: \let\G=\Gamma \let\D=\Delta \let\Th=\Theta\let\L=\Lambda \let\X=\Xi
126: \let\P=\Pi \let\Si=\Sigma \let\F=\Phi \let\Ps=\Psi
127: \let\O=\Omega \let\Y=\Upsilon
128:
129: \def\*{\vglue3truemm}
130: \let\0=\noindent
131: \def\otto{\,{\kern-1.truept\leftarrow\kern-5.truept\to\kern-1.truept}\,}
132: \def\media#1{{\langle#1\rangle}}
133: \def\defi{\,{\buildrel def\over=}\,}
134: \def\EE{{\cal E}}\def\NN{{\cal N}}\def\FF{{\cal F}}\def\CC{{\cal C}}
135: \def\V#1{{\bf#1}}\def\wt#1{{\widetilde#1}}
136: \def\tende#1{\,\vtop{\ialign{##\crcr\rightarrowfill\crcr
137: \noalign{\kern-1pt\nointerlineskip} \hskip3.pt${\scriptstyle
138: #1}$\hskip3.pt\crcr}}\,}
139: \def\T#1{{#1_{\kern-3pt\lower7pt\hbox{$\widetilde{}$}}\kern3pt}}
140: \def\W#1{#1_{\kern-3pt\lower7.5pt\hbox{$\widetilde{}$}}\kern2pt\,}
141: \def\lis#1{\overline#1}
142: \def\wt#1{\widetilde#1}\def\wh#1{\widehat#1}
143: \def\bra#1{{\langle#1}}\def\ket#1{{|#1\rangle}}
144: \def\brav#1{{\langle#1|}}
145: \def\AA{{\cal A}}\def\DD{{\cal D}}\def\BB{{\cal B}}\def\PP{{\cal P}}
146: %%%%%%%%%%%%%%%
147: \def\Im{{\rm Im}\,}
148: \def\be{\begin{equation}}\def\ee{\end{equation}}
149: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
150: \def\LL{{\cal L}}
151:
152: %%%%%%%%%%%%%%% Pacchetti %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
153: \usepackage{makeidx} % allows index generation
154: \usepackage{eqalignno}
155:
156: \usepackage{fancyhdr}\pagestyle{fancy}{}\fancyhead{}\fancyfoot{}
157: \headheight=13.6pt %opzione
158:
159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160: \renewcommand{\headrulewidth}{0pt}
161: \def\iniz{\setcounter{equation}{0}{
162: \rhead{\thepage}\lhead{{{{\diecibf\thesection:}\ \SEC}}}}}
163:
164: %\def\iniz{\setcounter{equation}{0}{%
165: %\ifodd\thepage{\rhead{\thepage}\lhead{\SEC}\message{Dispari}}%
166: %\else{\lhead{\thepage}\rhead{\SEC}\message{Pari}}\fi}\*}
167:
168: %%%%%%%%%%%%%%% Fine Pacchetti %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169:
170: \begin{document}
171: \maketitle
172:
173: \begin{abstract}{\it The Heat theorem reveals the second law of
174: equilibrium Thermodynamics (i.e.existence of Entropy) as a
175: manifestation of a general property of Hamiltonian Mechanics and of
176: the Ergodic Hypothesis, valid for $1$ as well as $10^{23}$ degrees
177: of freedom systems, {\it i.e.} for simple as well as very complex
178: systems, and reflecting the Hamiltonian nature of the microscopic
179: motion. In Nonequilibrium Thermodynamics theorems of comparable
180: generality do not seem to be available. Yet it is possible to find
181: general, model independent, properties valid even for simple chaotic
182: systems ({\it i.e.} the hyperbolic ones), which acquire special
183: interest for large systems: the Chaotic Hypothesis leads to the
184: Fluctuation Theorem which provides general properties of certain
185: very large fluctuations and reflects the time-reversal symmetry.
186: Implications on Fluids and Quantum systems are briefly hinted. The
187: physical meaning of the Chaotic Hypothesis, of SRB distributions and
188: of the Fluctuation Theorem is discussed in the context of their
189: interpretation and relevance in terms of Coarse Grained Partitions
190: of phase space. This review is written taking some care that each section
191: and appendix is readable either independently of the rest or with
192: only few cross references.}\end{abstract} \*
193: %$\Bf=\Ba=\Bb$.
194: \vfill\eject
195:
196: % {\baselineskip=10pt
197: \tableofcontents
198:
199: \kern3mm
200: {\hbox{\large\bf \kern0mm 19 References\kern93.3mm\pageref{secRef}}}
201: % }
202: \vfill\eject
203:
204: \def\SEC{Heat Theorem}
205: \section{The Heat Theorem}
206: \label{sec1}\iniz
207:
208: An important contribution of Boltzmann to Physics as well as to
209: research methods in Physics has been the {\it Heat Theorem}.
210:
211: Summarizing here an intellectual development, spanning about twenty
212: years of work, the {\it Heat Theorem} for systems of particles of
213: positions $\V q$ and momenta $\V p$, whose dynamics is modeled by a
214: Hamiltonian of the form $H=K(\V p)+ W(\V q)$, $K=\frac1{2m}\V p^2$, can
215: be formulated as follows
216: \*
217:
218: \0{\bf Heat Theorem:} {\it In a isolated mechanical system,
219: %in thermodynamic equilibrium state the
220: time averages $\media{F}$ of the observables,
221: {\it i.e.} of functions $F$ on phase space, are computable as their
222: integrals with respect to probability distributions $\m_\Ba$ which
223: depend on the control parameters $\Ba$ determining the states. It is
224: possible to find four observables, whose averages can be called
225: $U,V,T,p$, depending on $\Ba$, so that an infinitesimal change $d\Ba$
226: implies variations $dU,dV$ of $U,V$ so related that
227:
228: \be \frac{dU+p\, dV}{T}=\,\hbox{``exact''}\,\defi dS\label{1.1}\ee
229: %
230: where $p=\media{-\partial_V W}$ and $V$ is a(ny) parameter on which $W$
231: depends, and $U,T$ are the average total energy and the average
232: total kinetic energy.
233: \\
234: When the system is large and $V$ is the volume available to
235: the particles the quantity $p$ can be shown to have the interpretation
236: of physical ``pressure'' on the walls of the available volume.}
237: \*
238:
239: \0{\it Remarks:} (a) Identification of $T$ with the average kinetic energy
240: had been for Boltzmann a starting point, assumed a priori, from the
241: works of Kr\"onig and Clausius of a few years earlier (all apparently
242: unaware, as everybody else, of the works of Bernoulli, Herapath,
243: Waterstone, \cite{Br03}).
244: \\
245: %
246: (b) Connection with observations is made by identifying curves in
247: parameter space, $t\to\Ba(t)$, with {\it reversible processes}. And in
248: an infinitesimal process, defined by a line element $d\Ba$, the
249: quantity $pdV$ is identified with the work the system performs, $dU$
250: with the energy variation and $dQ=dU+pdV$ as the heat absorbed. Then
251: relation Eq.(\ref{1.1}) implies that Carnot machines have the highest
252: efficiency. The latter is one of the forms of the second law, which
253: leads to the existence of entropy as a function of state in macroscopic
254: Thermodynamics, \cite{Ze68}.
255: \\
256: %
257: (c) Eq.(\ref{1.1}), combined with the (independent) assumption that
258: heat extracted at a fixed temperature cannot be fully transformed into
259: work, implies that in any process $\frac{dQ}T\le dS$. Hence in
260: isolated systems changing equilibrium state cannot make entropy
261: decrease, or in colorful language the {\it entropy of the Universe
262: cannot decrease}, \cite[p. I-44-12]{Fe63}. Actually by suitably defining
263: what is meant by irreversible process it is possible to reach the
264: conclusion that, unless the change of equilibrium state is achieved
265: via a reversible process, the entropy of an isolated system does
266: increase strictly, \cite{Ze68}. Conceptually, however, this is an
267: addition to the second law, \cite[p. I-44-13]{Fe63}. \*
268:
269: Examples of control parameters are simply $U,V$, or $T,V$, or $p,V$.
270: The theorem holds under some hypotheses which evolved from
271: \*
272:
273: \0(a) all motions are periodic (1866)
274: %
275: \\
276: %
277: (b) aperiodic motions can be considered periodic with infinite
278: period (!), \cite{Bo66}.
279: %
280: \\
281: %
282: (c) motion visits all phase space of given total energy: in modern
283: terminology this is the {\it ergodic hypothesis} (1868-1884),
284: \cite{Bo884}. \*
285:
286: The guiding idea was that Eq.(\ref{1.1}) would be true for all systems
287: described by a Hamiltonian $H=K+W$: {\it no matter whether having few
288: or many degrees of freedom}, as long as the ergodic hypothesis could
289: be supposed true.
290:
291: In other words Eq.(\ref{1.1}) should be considered as a consequence of
292: the Hamiltonian nature of motions: it is true for all systems
293: whether with one degree of freedom (as in the 1866 paper by Boltzmann)
294: or with $10^{19}$ degrees of freedom (as in the 1884 paper by Boltzmann).
295:
296: It is, in a sense, a property of the particular Hamiltonian structure
297: of Newton's equations (Hamiltonian given as sum of kinetc plus
298: potential energy with kinetic energy equal to $\sum_i\frac12{\V
299: p}_i^2$ and potential energy purely positional). True for all
300: (ergodic) systems: trivial for $1$ degree of freedom, a surprising
301: curiosity for few degrees and an important law of Nature for $10^{19}$
302: degrees of freedom (as in $1\,cm^3$ of ${\rm H_2}$).
303:
304: The aspect of Boltzmann's approach that will be retained here is that
305: some universal laws merely reflect basic properties of the equations
306: of motion which may have deep consequences in large systems: the roots
307: of the second Law can be found, \cite{Bo66}, in the simple properties
308: of the pendulum motion.
309:
310: Realizing the mechanical meaning of the second law induced the
311: birth of the theory of ensembles, developed by Boltzmann between
312: 1871 (as recognized by Gibbs in the introduction to his treatise) and
313: 1884, hence of Statistical Mechanics.
314:
315: Another example of the kind are the reciprocal relations of Onsager,
316: which reflect time reversal symmetry of the Hamiltonian systems
317: considered above. Reciprocity relations are a first step towards
318: understanding non equilibrium properties. They impose strong
319: constraints on transport coefficients, {\it i.e.} on the $\V
320: E$-derivatives of various average currents induced by external
321: forces of intensities $\V E=(E_1,\ldots,E_n)$, which disturb the system
322: from an equilibrium state into a new {\it stationary
323: state}. The derivation leads to the quantitative form of reciprocity
324: which is expressed by the ``Fluctuation-Dissipation Theorems'', {\it
325: i.e.} by the Green-Kubo formulae, expressing the transport
326: coefficient of a current in terms of the mean square fluctuations of
327: its long time averages.
328:
329: In the above Boltzmann's papers (as well as in several other of his
330: works) Thermodynamics is derived on the assumption that motions are
331: periodic, hence very regular: see the above mentioned ergodic
332: hypothesis. Nevertheless heat is commonly regarded as associated with
333: the chaotic motions of molecules and thermal phenomena are associated
334: with fluctuations due to chaotic motions at molecular level. A theme
335: that is pursued in this paper it to investigate how to reconcile
336: opposites like order and chaos within a unified approach so general to
337: cover not only equilibrium Statistical Mechanics, but many aspects of
338: nonequilibrium stationary states. An overview is in the first thiteen
339: sections, while the appendices enter into technical details, still
340: keeping at a heuristic level in discusing a matter that is often given
341: little conideration by Physicists because of its widespread reputation
342: of being just abstract Mathematics: hopefully this will help to
343: divulge a theory which is not only simple conceptually nut it seems
344: promising of further developments.
345:
346: The above comment is meant also to explain the meaning of the title of
347: this paper.
348: \*
349:
350: \def\SEC{Time Reversal Symmetry}
351: \section{Time Reversal Symmetry}
352: \label{sec2}\iniz
353: \*
354:
355: In a way transport coefficients are still equilibrium properties and
356: nothing is implied by reciprocity when $\V E$ is strictly $\ne\V 0$.
357:
358: It is certainly interesting to investigate whether time reversal has
359: important implications in systems which are really out of equilibrium,
360: {\it i.e.} subject to non conservative forces which generate currents
361: (transporting mass, or charge, or heat or several of such
362: quantities).
363:
364: There have been many attempts in this direction: it is important to
365: quote the reference \cite{BK81a} which summarizes a series of works by
366: a Russian school and completes them. In this paper an extension of the
367: Fluctuation-Dissipation theorem, as a reflection of time reversal, is
368: presented, deriving relations which, after having been further
369: developed, have become known as ``work theorems'' and/or ``transient
370: fluctuation theorems'' for transformations of systems out of
371: equilibrium, \cite{Ja97,Ja99,ES94,Cr99,CG99,Ga06b}.
372:
373: For definiteness it is worth recalling that a dynamical system with
374: equations $\dot x= f(x)$ in phase space, whose motions will be given
375: by maps $t\to S_t x$, is called ``reversible'' if there is a
376: smooth ({\it i.e.} continuously differentiable) isometry $I$ of phase space,
377: anticommuting with $S_t$ and involutory, {\it i.e.}
378:
379: \be I S_t=S_{-t}I,\qquad I^2=1\label{2.1}\ee
380: %
381: Usually, if $x=(\V p,\V q)$, time reversal is simply $I(\V p,\V q)=(-\V
382: p,\V q)$.
383:
384: The main difficulty in studying nonequilibrium statistical Mechanics
385: is that, after realizing that one should first understand the
386: properties of stationary states, considered as natural extensions of the
387: equilibrium states, it becomes clear that the microscopic description
388: {\it cannot be Hamiltonian}.
389:
390: This is because a current arising from the action of a nonconservative
391: force continuously generates ``heat'' in the system. Heat has to be
392: taken out to allow reaching a steady state. This is empirically done
393: by putting the system in contact with one or more thermostats. In
394: models, thermostats are just forces which act performing work
395: balancing, at least in average, that produced by the external
396: forces, {\it i.e.} they ``model heat extraction''.
397:
398: It is not obvious how to model a thermostat; and any thermostat model
399: is bound to be considered ``unphysical'' in some respects. This is not
400: surprising, but it is expected that most models introduced to describe
401: a given physical phenomenon should be ``equivalent''.
402:
403: Sometimes it is claimed that the only physically meaningful
404: thermostats for nonequilibrium systems (in stationary states) are made
405: by infinite ($3$-dimensional) systems which, asymptotically at
406: infinity, are in statistical equilibrium. In the latter cases it is
407: not even necessary to introduce {\it ad hoc} forces to remove the
408: heat: {\it motion remains Hamiltonian} and heat flows towards
409: infinity.
410:
411: Although the latter is certainly a good and interesting model, as
412: underlined already in \cite{FV63}, it should be stressed that it is
413: mathematically intractable unless the infinite systems are
414: ``free''. {\it i.e.} without internal interaction other than linear,
415: \cite{FV63,HL73b,EPR99,APJP06,HI05}.
416:
417: And one can hardly consider such assumption more physical than the one
418: of finite thermostats. Furthermore it is not really clear whether a linear
419: external dynamics can be faithful to Physics, as shown by the simple
420: one dimensional XY-models, see \cite{ABGM72} where a linear thermostat
421: dynamics with a single temperature leads a system to a stationary
422: state, as expected, but the state is not a Gibbs state (at any
423: temperature). The method followed in \cite{ABGM72}, based on
424: \cite{BD69}, can be used to illustrate some problems which can arise
425: when thermostats are classical free systems, see Appendix A4.
426:
427: \def\SEC{Point of view}
428: \section{Point of view}
429: \label{sec3}\iniz
430:
431: The restriction to finite thermostats, followed here, is not chosen
432: because infinite thermostats should be considered
433: unphysical, but rather because it is a fact that the recent progress in
434: nonequilibrium theory can be traced to
435: \*
436:
437: \0(a) the realization of the interest of restricting attention
438: to {\it stationary states}, or {\it
439: steady states}, reached under forcing (rather than discussing approach
440: to equilibrium, or to stationarity).
441:
442: \0(b) the simulations on steady states performed in the 80's after the
443: essential role played by {\it finite thermostats} was fully realized.
444: \*
445:
446: Therefore investigating finite thermostat models is still particularly
447: important. This makes in my view interesting to confine attention on
448: them and to review their conceptual role in the developments that took
449: place in the last thirty years or so.
450:
451: Finite thermostats can be modeled in several ways: but in constructing
452: models it is desirable that the models keep as many features as
453: possible of the dynamics of the infinite thermostats. As realized in
454: \cite[p.452]{BK81a} it is certainly important to maintain the {\it
455: time reversibility}. Time reversibility expressed by Eq. (\ref{2.1}), {\it
456: i.e.} existence of a smooth conjugation between past and future,
457: is a fundamental symmetry of nature which (replaced by TCP) even
458: ``survives'' the so called time reversal violation; hence it is desirable
459: that it is saved in models. An example will be discussed later.
460: \*
461:
462: \0{\it Comment:} (1) The second law of equilibrium Thermodynamics,
463: stating existence of the state function entropy, can be derived
464: without reference to the microscopic dynamics by assuming that heat
465: absorbed at a single temperature cannot be cyclically converted into
466: work, \cite{Ze68}. In statistical Mechanics equilibrium, states are
467: identified with probability distributions on phase space: they depend
468: on control parameters (usually two, for instance energy and volume)
469: and processes are identified with sequences of equilibrium states,
470: {\it i.e.} as curves in the parameters space interpreted as {\it
471: reversible processes}. The problem of how the situation, in which
472: averages are represented by a probability distribution, develops
473: starting from an initial configuration {\it is not part of the
474: equilibrium theory}. In this context the second law arises as a
475: theorem in Mechanics (subject to asssumptions) and, again, just says
476: that entropy exists (the heat theorem).
477: %
478: \\
479: %
480: (2) As noted in Sec.\ref{sec1}, if the scope of the theory is enlarged
481: admitting processes that cannot be represented as sequences of
482: equilibria, called ``irreversible processes'', then the postulate of
483: impossibility to convert heat into work extracting it from a single
484: thermostat implies, again without involving microscopic dynamics, the
485: inequality often stated as ``the entropy of the Universe'' cannot
486: decrease in passing from an equilibrium state to another. And, after
487: properly defining what is meant by irreversible process \cite{Ze68},
488: actually strictly increases if in the transformation an irreversible
489: process is involved; however perhaps it is best to acknowledge
490: explicitly that such a strict increase is a further assumption,
491: \cite[p. I-44-13]{Fe63} leaving aside a lengthy, \cite{Ze68}, and
492: possibly not exhaustive analysis of how in detail an irreversible
493: transformation looks like. Also this second statement, under suitable
494: assumptions, can become a theorem in Mechanics, \cite{Le93,Ga01}, but
495: here this will not be discussed.
496: %
497: \\
498: %
499: (3) Therefore studying macroscopic properties for systems out of
500: equilibrium can be divided into an ``easier'' problem, which is the
501: proper generalization of equilibrium statistical Mechanics: namely
502: studying stationary states identified with corresponding probability
503: distributions yielding, by integration, the average values of the few
504: observables of relevance. And the problem of approach to a stationary
505: state which is of course more difficult. The recent progress in
506: nonequilibrium has been spurred by restricting research to the easier
507: problem.
508:
509: \def\SEC{Chaotic Hypothesis}
510: \section{The Chaotic Hypothesis (CH)}
511: \label{sec4}\iniz
512:
513: Following Boltzmann and Onsager we can ask whether there are general
514: relations holding among time averages of selected observables and for
515: all systems that can be modeled by time reversible mechanical
516: equations $\dot x=f(x)$.
517:
518: The difficulty is that in presence of dissipation it is by no means
519: clear which is the probability distribution $\m_\Ba$ which provides
520: the average values of observables, at given control parameters $\Ba$.
521:
522: In finite thermostat models dissipation is manifested by the
523: nonvanishing of the divergence, $\s(x)\defi-\sum \partial_{x_i}
524: f_i(x)$, of the equations of motion and of its time average $\s_+$.
525:
526: If $\s_+>0$%
527: \footnote{\nota As intuition suggests $\s_+$ cannot be $<0$,
528: \cite{Ru97}, when motion takes place in a bounded region of phase
529: space, as it is supposed here.}, it is not possible that the
530: distributions $\m_\Ba$ be of the form $\r_\Ba(x)dx$, ``absolutely
531: continuous with respect to the phase space volume'': since volume
532: contracts, the probability distributions that, by integration, provide
533: the averages of the observables must be concentrated on sets,
534: ``attractors'', of $0$ volume in phase space.
535:
536: This means that there is no obvious substitute of the ergodic
537: hypothesis: which, however, was essential in equilibrium statistical
538: Mechanics to indicate that the ``statistics'' $\m_\Ba$, {\it i.e.} the
539: distribution $\m_\Ba$ such that
540:
541: \be \lim_{T\to\infty}\frac1T\int_0^T F(S_tx)dt=\int \m_\Ba(dy)
542: F(y)\label{4.1}\ee
543: %
544: for all $x$ except a set of zero volume, exists and is given by the
545: Liouville volume (appropriately normalized to $1$) on the surfaces of
546: given energy $U$ (which is therefore one of the parameters $\Ba$ on
547: which the averages depend).\footnote{\nota By Liouville volume we mean
548: the measure $\d(K(\V p)+ W(\V q)-U)d\V p d\V q$, on the manifold of
549: constant energy or, in dissipative cases discussed later, the measure
550: $d\V p d\V q$.\label{2}\label{footnote2}}%%%%for page ref volume
551:
552: It is well known that identifying $\m_\Ba$ with the Liouville volume
553: does not allow us to derive the values of the averages (aside from a
554: few very simple cases, like the free gas): but it allows us to write
555: the averages as explicit integrals, \cite{Ga00}, which are well suited to
556: deduce relations holding between certain averages, like the second law
557: Eq.(\ref{1.1}) or Onsager reciprocity and the more general Fluctuation
558: Dissipation Theorems.
559:
560: The problem of finding a useful representation of the statistics of
561: the stationary states in systems which are not in equilibrium arose in
562: the more restricted context of fluid Mechanics earlier than in
563: statistical Mechanics. And through a critique of earlier attempts,
564: \cite{Ru78b}, in 1973 Ruelle proposed that one should take advantage of
565: the empirical fact that motions of turbulent systems are ``chaotic''
566: and suppose that their mathematical model should be a ``hyperbolic
567: system'', in the same spirit in which the ergodic hypothesis should be
568: regarded: namely {\it while one would be very happy to prove
569: ergodicity because it would justify the use of Gibbs' microcanonical
570: ensemble, real systems perhaps are not ergodic but behave nevertheless
571: in much the same way and are well described by Gibbs' ensemble...},
572: \cite{Ru73}.
573:
574: The idea has been extended in \cite{GC95,Ga00} to nonequilibrium statistical
575: Mechanics in the form
576:
577: \* \0{\bf Chaotic hypothesis (CH):} {\it Motions on the attracting set
578: of a chaotic system can be regarded as motions of a smooth transitive
579: hyperbolic system.\footnote{\nota Transitive means ``having a dense
580: orbit''. Note that here this is a property of the attracting set,
581: which is often not at all dense in the full phase space. Such systems
582: are also called ``Anosov systems''.}} \*
583:
584: The hypothesis was formulated to explain the result of the experiment
585: in \cite{ECM93}. In \cite{GC95} it was remarked that the CH could be
586: adequate for the purpose.
587:
588: \*
589: \def\SEC{``Free'' implications of CH}
590: \section{``Free'' implications of the Chaotic Hypothesis}
591: \label{sec5}\iniz
592:
593: Smooth transitive hyperbolic systems share, independently of the number
594: of degrees of freedom, remarkable properties, \cite{GBG04}. \*
595:
596: \0(1) their motions can be considered paradigmatic chaotic evolutions,
597: whose theory is, nevertheless, very well understood to the point
598: that they can play for chaotic motions a role alike to the one
599: played by harmonic oscillators for ordered motions, \cite{Ga95a}.
600:
601: \0(2) there is a {\it unique} distribution $\m$ on phase space such that
602:
603: \be \lim_{\t\to\infty}\frac1\t\int_0^\t F(S_tx)dt=\int \m(dy) F(y)
604: \label{5.1}\ee
605: %
606: for all smooth $F$ and for all but a zero volume set of initial data
607: $x$, \cite{Si72,BR75,Ga00,GBG04}, see Appendix A1. The distribution
608: $\m$ is called the {\it SRB probability distribution}, see Appendix A2.
609:
610: \0(3) averages satisfy a {\it large deviations rule}:
611: {\it i.e.} if the point $x$ in $f=\frac1\t\int_0^\t F(S_tx)\,dt$
612: is sampled with distribution $\m$, then
613:
614: \be \lim_{\t\to\infty}
615: \frac1\t\log Prob_\m(f\in\D)= \max_{f\in\D} \z_F(f)\label{5.2}\ee
616: %
617: is an asymptotic value that controls the probability that the finite
618: time average of $F$ falls in an interval $\D=[u,v], \,u<v$, subset of
619: the interval $(a_F,b_F)$ of definition of $\z_F$. In the interval of
620: definition $\z_F(f)$ is convex and analytic in $f$, \cite{Si72,Si77}.
621: Outside $[a_F,b_F]$ the function $\z_F(f)$ can be defined to have
622: value $-\infty$ (which means that values of $f$ in intervals outside
623: $[a_F,b_F]$ can possibly be observed only with a probability tending
624: to $0$ faster than exponentially), \cite{Si72,Si77}.
625:
626: \0(4) A more precise form of Eq.(\ref{5.2}) yields also the rate at
627: which the limit is reached: $Prob_\m(f\in\D)= e^{\t\,\max_{f\in\D}
628: \z_F(f)+ O(1)}$ with $O(1)$ bounded uniformly in $\t$, at fixed
629: distance of $\D$ from the extremes $a_F,b_F$. This is ofteen written
630: in a not very precise but mnemocnically convenient form, as long as
631: its real meaning is kept in mind, as
632:
633: \be P_\m(f)=e^{\t\,\z_F(f)+O(1)}\label{5.3}\ee
634:
635: \0(5) The fluctuations described by (\ref{5.2}) are very large
636: fluctuations as they have size of order $\t$ rather than
637: $O(\sqrt{\t})$: in fact if the maximum of $\z_F(f)$ is at a point
638: $f_0\in(a_F,b_F)$ and is a nondegenerate quadratic maximum, then
639: Eq. (\ref{5.2}) implies that $\sqrt{\t}(f-f_0)$ has an asymptotically
640: Gaussian distribution. This means that the motion can be regarded to be so
641: chaotic that the values of $F(S_tx)$ are independent enough so that
642: the finite time average deviations from the mean value $f_0$ are
643: Gaussian on the scale of $\sqrt\t$.
644:
645: \0(6) A natural extension to (\ref{5.2}) in which several observables
646: $F_1,\ldots,F_n$ are simultaneously considered is obtained by defining
647: $f_i=\frac1\t\int_0^\t F_i(S_tx)dt$. Then there exists a convex closed
648: set $C\subset {\cal R} ^n$ and function $\z_{\V F}(\V f)$ analytic in $\V
649: f=(f_1,\ldots,f_n)$ in the interior of $C$ and, given an open set
650: $\BD\subset C$,
651:
652: \be \lim_{\t\to\infty} \frac1\t\log Prob_\m(\V f\in\BD)= \max_{\V
653: f\in\BD}\,\,\z_{\V F}(\V f)\label{5.4}\ee
654: %
655: and $\z_{\V F}(\V f)$ could be defined as $-\infty$ outside $C$, with
656: the meaning mentioned in remark (2). If the function $\z_{\V F}(\V f)$
657: attains its maximum in a point $\V f_0$ in the interior of $C$ and the
658: maximum is quadratic and nondegenerate, then the {\it joint
659: fluctuations} of $\Bf=\sqrt\t(\V f-\V f_0)$ are asymptotically
660: Gaussian, which means that have a probability density
661: $\frac1{\sqrt{\p^n\det \DD}} e^{-\frac12 (\Bf\cdot \DD^{-1}\Bf)}$ with
662: $\DD$ a positive definite $n\times n$ matrix.
663:
664: \0(7) The probability distribution $\m$ depends on the control
665: parameters $\Ba$ of the initial data and therefore as $\Ba$ varies
666: one obtains a collection of probability distributions: this leads
667: to a natural {\it extension of the ensembles of equilibrium
668: statistical Mechanics}, \cite{Ga00}.
669:
670: \0(8) The most remarkable property, root of all the above, is that the
671: SRB probability distribution $\m$, can be given a concrete formal
672: representation, in spite of being a distribution concentrated on a
673: set of zero volume, \cite{Si72,Si77}, see Appendix A1,A2. This raises
674: hopes to use it to derive general relations between averages of
675: observables. As in equilibrium, the averages with respect to $\m$
676: are destined to remain not computable except, possibly, under
677: approximations (aside very few exactly soluble cases): their
678: formal expressions could nevertheless be used to establish general
679: mutual relations and properties.
680:
681: \0(9) Given the importance of the existence and representability of
682: the SRB distribution, Appendix A1,A2 will be entirely devoted to
683: the formulation (A1) and to the physical interpretation of the
684: derivation of its expression: this could be useful for readers who
685: want to understand the technical aspects of what follows, because
686: some may find not satisfactory skipping the technical details even
687: at a heuristic level. The aim of the non technical discussion that
688: follows, preceding the appendices, is to make it worth to invest
689: some time on the technical details.
690:
691: \0(10) Applied to a system in equilibrium the CH implies the ergodic
692: hypothesis so that it is a genuine extension of the latter and any
693: results that follow from it will be necessarily compatible with
694: those of equilibrium statistical Mechanics, \cite{Ga00}.
695:
696: \0(11) For very simple systems the distribution $\m$ can be constructed
697: explicitly and time averages of some observables computed. The
698: systems are the discrete time evolutions corresponding to linear
699: hyperbolic maps of tori, \cite{GBG04}, or the continuous time geodesic
700: motion on a surface of constant negative curvature. The latter
701: systems are rigorously hyperbolic and the SRB distribution can be
702: effectively computed for them {\it as well as for their small
703: perturbations}.
704:
705: \0(12) A frequent remark about the chaotic hypothesis is that it does
706: not seem to keep the right viewpoint on nonequilibrium
707: Thermodynamics. It should be stressed that the hypothesis is
708: analogous to the ergodic hypothesis, which ({\it as well known})
709: cannot be taken as the foundation of equilibrium statistical
710: Mechanics, even though it leads to the correct Maxwell Boltzmann
711: statistics, because the latter ``holds for other reasons''.
712: Namely it holds because in most of phase space (measuring sizes by
713: the Liouville measure) the few interesting macroscopic observables
714: have the same value, \cite{Th74}, see also \cite{Le93}. \*
715:
716: \def\SEC{Paradigms}
717: \section{Paradigms of Statistical Mechanics and CH}
718: \label{sec6}\iniz
719:
720: In relation to the last comment is useful to go back to the Heat
721: Theorem of Sec.1 and to a closer examination of the basic paper of
722: Boltzmann \cite{Bo884}, in which the theory of equilibrium
723: ensembles is developed and may offer arguments for further
724: meditation. The paper starts by illustrating an important, and
725: today almost forgotten, remark by Helmoltz showing that very
726: simple systems (``monocyclic systems'') can be used to construct
727: mechanical models of Thermodynamics: and the example chosen by
728: Boltzmann is {\it really extreme by all standards}.
729:
730: He shows that the motion of a {\it Saturn ring} of mass $m$ on
731: Keplerian orbits of major semiaxis $a$ in a gravitational field of
732: strength $g$ can be used to build a model of Thermodynamics. In
733: the sense that one can call \
734: \*
735: \0``volume'' $V$ the gravitational
736: constant $g$,
737: \\
738: ``temperature'' $T$ the average kinetic energy,
739: \\
740: ``energy'' $U$ the energy and
741: \\
742: ``pressure'' $p$ the average potential
743: energy $m k a^{-1}$,
744: \*
745: \0then one infers that by varying, {\it at fixed
746: eccentricity}, the parameters $U,V$ the relation
747: $(dU+pdV)/T=\,\hbox{\it exact}$ holds. Clearly this {\it could}
748: be regarded as a curiosity, see \cite[Appendix 1.A1, Appendix 9.A3]{Ga00}.
749:
750: However Boltzmann (following Helmoltz?\footnote{\nota The relation
751: between the two on this subject should be more
752: studied. Boltzmann's paper of 1884, \cite{Bo884}, is a natural
753: follow up and completion of his earlier work \cite{Bo871b} which
754: followed \cite{Bo868,Bo66}. It seems that the four extremely long
755: papers by Helmoltz, also dated 1884, \cite{He884a,He884b}, might
756: have at most just stimulated Boltzmann to revisit his earlier
757: works and led him achieve the completion of the mechanical
758: explanation of the second law. Certainly Boltzmann attributes a
759: strong credit to Helmoltz, and one wonders if this might be partly
760: due to the failed project that Boltzmann had to move to Berlin
761: under the auspices of Helmoltz.}) took it seriously and proceeded
762: to infer that under the ergodic hypothesis {\it any system} small
763: or large provides us with a model of Thermodynamics (being
764: ``monocyclic'' in the sense of Helmoltz): for instance he showed
765: that the canonical ensemble verifies exactly the second law of
766: equilibrium Thermodynamics (in the form $(dU+p\,dV)/T=\, \hbox{\it
767: exact}$) {\it without any need to take thermodynamic limits},
768: \cite{Bo884}, \cite{Ga00}. The same could be said of the
769: microcanonical ensemble (here, however, he had to change
770: ``slightly'' the definition of heat to make things work without
771: finite size corrections).
772:
773: He realized that the Ergodic Hypothesis could not possibly account
774: for the correctness of the canonical (or microcanonical)
775: ensembles; this is clear at least from his (later) paper in
776: response to Zermelo's criticism, \cite{Bo96}. Nor it could account
777: for the observed time scales of approach to equilibrium.
778: Nevertheless he called the theorem he had proved the {\it heat
779: theorem} and never seemed to doubt that it provided evidence for
780: the correctness of the use of the equilibrium ensembles for
781: equilibrium statistical Mechanics.
782:
783: Hence there are two points to consider: first certain
784: relations among mechanical quantities {\it hold no matter how
785: large} is the size of the system and, secondly, they can be seen
786: and tested not only in small systems, by direct measurements, but
787: even in large systems, because in large systems such mechanical
788: quantities acquire a macroscopic thermodynamic meaning and their
789: relations are ``typical'' {\it i.e.} they hold in most of phase space.
790:
791: The first point has a close analogy in that the consequences of
792: the Chaotic Hypothesis stem from the properties of small dimension
793: hyperbolic systems (the best understood) which play here the role
794: of Helmoltz' monocyclic systems of which Boltzmann's Saturn ring
795: (\cite{Bo884}) is a special case. They are remarkable
796: consequences because they provide us with {\it parameter free
797: relations} (namely the Fluctuation Theorem, to be discussed below,
798: and its consequences): but clearly it cannot be hoped that a
799: theory of nonequilibrium statistical Mechanics be founded solely
800: upon them, by the same reasons why the validity of the second law
801: for monocyclic systems had in principle no reason to imply the
802: theory of ensembles.
803:
804: Thus what is missing are arguments similar to those used by
805: Boltzmann to justify the use of ensembles, {\it independently} of
806: the ergodic hypothesis: an hypothesis which in the end may appear
807: (and still does appear to many) as having only suggested them ``by
808: accident''. The missing arguments should justify the CH on the
809: basis of an extreme likelihood of its predictions in systems that
810: are very large and that may be not hyperbolic in the mathematical
811: sense. I see no reason, now, why this should prove impossible
812: {\it a priori} or in the future. See Sect.\ref{sec12} for some of
813: the difficulties that can be met in experiments testing the CH
814: through its consequence discussed in Sec.\ref{sec7}.
815:
816: \* In the meantime it seems interesting to take the same philosophical
817: viewpoint adopted by Boltzmann: not to consider a chance that {\it
818: all} chaotic systems share some selected, and remarkable,
819: properties and try to see if such properties help us achieving a
820: better understanding of nonequilibrium. After all it seems that
821: Boltzmann himself took a rather long time to realize the interplay
822: of the just mentioned two basic points behind the equilibrium
823: ensembles and to propose a solution harmonizing them. ``All it
824: remains to do'' is to explore if the hypothesis has implications
825: more interesting or deeper than the few known and presented in the
826: following.
827:
828: \def\SEC{Fluctuation Theorem}
829: \section{The Fluctuation Theorem (FT)}
830: \label{sec7}\iniz
831:
832: The idea of looking into time reversibility to explain the
833: experimental results of \cite{ECM93} is clearly expressed in the same
834: paper. The CH allows us to use effectively time reversal symmetry
835: to obtain what has been called in \cite{GC95,Ga95b,GC95b} the ``{\it
836: Fluctuation Theorem}''. In fact a simple property holds for all
837: transitive hyperbolic systems which admit a time reversal symmetry.
838:
839: The property deals with the key observable $\s(x)$, which is the above
840: introduced divergence of the equations of motion, or ``phase space
841: contraction rate''. Assuming the average phase space contraction to be
842: positive, $\s_+>0$, let $p=\frac1\t\int_0^\t \frac{\s(S_t x)}{\s_+}dt$
843: be the ``dimensionless phase space contraction''; let $\z(p)$ be the large
844: deviation rate function introduced in \S\ref{sec5}, see
845: Eq.(\ref{5.2}), for $F(x)=\frac{\s(x)}{\s_+}$. By time reversal
846: symmetry the interval of analyticity of $\z(p)$ is centered at the
847: origin and will be denoted $(-p^*,p^*)$; furthermore $p^*\ge1$,
848: because the average of $p$ is $1$. Then, \cite{GC95},
849:
850: \*
851: \0{\bf Fluctuation Theorem (FT):}
852: {\it The probabilities of the large deviations of $p$ satisfy, for all
853: transitive time reversible hyperbolic systems,
854:
855: \be \z(-p)=\z(p)-p\s_+\label{7.1}\ee
856: %
857: for all $|p|<p^*$: this will be called a ``fluctuation relation'',
858: (FR). } \*
859:
860: \0{\it Remarks:}
861: %
862: \\
863: %
864: (1) In terms of the notation in Eq.(\ref{5.3}) the FT is
865:
866: \be\frac{P_\t(p)}{P_\t(-p)}=e^{\,p\,\s_+\,\t +O(1)}\label{7.2}\ee
867: %
868: which is the form in which it is often written.
869: %
870: \\
871: %
872: (2) The theorem has been developed, in \cite{GC95}, to understand the
873: results of a simulation, \cite{ECM93}, whose Authors had correctly
874: pointed out that the SRB distribution together with the time
875: reversibility could possibly explain the observations.
876: %
877: \\
878: %
879: (3) Unfortunately the same name, introduced in \cite{GC95,Ga95b,GC95b}
880: where FT has been proved, has been {\it subsequently} picked up and
881: attributed to other statements, superficially related to the above
882: FT. Enormous confusion ensued (and sometimes even errors), see
883: \cite{CG99,GC04,GZG05}. A more appropriate name for such other, and
884: different, statements has been suggested to be ``transient fluctuation
885: theorems''. The above FT should be distinguished also from the
886: results in \cite{BK81a} which were the first {\it transient}
887: fluctuations results, later extended and successfully applied,
888: see \cite{Ja97,Ja99}. It is claimed that the difference between the
889: above FT and the transient statements is just an exchange of limits:
890: the point is that it is a nontrivial one, see counterexamples in
891: \cite{CG99}, and assumptions are needed, which have a physical
892: meaning; the CH is the simplest.
893: %
894: \\
895: %
896: (4) The FT theorem has been proved first for discrete time evolutions,
897: {\it i.e.} for maps: in this case the averages over time are
898: expressed by sums rather than by integrals. Hyperbolic maps are
899: simpler to study than the corresponding continuous time systems, which
900: we consider here, because smooth hyperbolic maps do not have a trivial
901: Lyapunov exponent (the vanishing one associated with the phase space
902: flow direction); but the techniques to extend the analysis to
903: continuous time systems are the same as those developed in \cite{Ge98}
904: for proving the FT for hyperbolic flows and in this review I shall not
905: distinguish between the two kinds of evolutions since the properties
906: considered here do not really differ in the two cases.
907: %
908: \\
909: %
910: (5) The condition $\s_+>0$, {\it i.e.} dissipativity, is {\it
911: essential} even to define $p$ itself. When the forcing intensity $E$
912: vanishes also $\s_+\to0$ and the FR loses meaning because $p$
913: does. Neverheless by appropriately dividing both sides of
914: Eq. (\ref{7.1}) by $\s_+$, and then taking the limit, a nontrivial
915: limit can be found and it can be shown, at least heuristically, to
916: give the Green-Kubo relation for the ``current''
917: $J\defi\media{\frac{\partial\s}{\partial E}}_{\m}=\media{j}_\m$,
918: \cite{Ga96a,Ga00}, generated by the forcing, namely
919:
920: \be \frac{d J}{d E}\Big|_{E=0}=\frac12\int_{-\infty}^\infty \media{j(S_\t
921: x)\,j(x)}_{E=0}dt\label{7.3}\ee
922: %
923: which is a general Fluctuation-Dissipa\-tion theorem.
924: %
925: \\
926: %
927: (6) The necessity of a bound $p^*$ in FT has attracted undue
928: attention: it is {\it obvious} that it is there since $\s(x)$ is
929: bounded, if CH holds. It also true that the role of $p^*$ is
930: discussed in the paper \cite{Ga95b}, which is a formal and {\it
931: contemporary} version of the earlier \cite{GC95} and of part of
932: the later \cite{GC95b} written for a different audience in mind.
933: %
934: \\
935: %
936: It is therefore surprising that this is sometimes ignored in the
937: literature and the original papers are faulted for not mentioning
938: this (obvious) point, which in any event is fully discussed in
939: \cite{Ga95b}. A proof which also discusses $p^*$ is in
940: \cite{Ru99}. It is also obvious that for $p\ge p^*$ the function
941: $\z(p)$ can be naturally set to be $-\infty$, as commented in
942: remark (6) to the CH in Sec.\ref{sec4}, and for this reason
943: Eq. (\ref{7.1}) is often written without any restiction on $p$.
944: This is another point whose misunderstanding has led to
945: errors. For readers familiar with statistical Mechanics there is
946: nothing misterious about $p^*$. It is analogous the ``close
947: packing density'' in systems with hard cores: it is clear that
948: there is a well defined maximum density but its value is not
949: always explicitly computable; and for hiher density many
950: thermodynamic functions may be considered defined but as having an
951: infinite value. \*
952:
953: \0{\bf Corollary:} {\cite{Ga97,Ga00},\it Under the same assumptions of FT,
954: if $F_1=\frac{\s(x)}{\s_+}$, $F_2,\ldots, F_n$ are $n$ observables of
955: parity $\e_i=\pm$ under time reversal, $F_i(Ix)=\e_i F_i(x)$, the
956: large deviations rate $\z_{\V F}(\V f)$, defined in Eq. (\ref{5.4}),
957: satisfies
958:
959: \be \z_{\V F}(\V f^*)=\z_{\V F}(\V f)-\s_+ f_1\label{7.4}\ee
960: %
961: where $\V f^*=(- f_1,\e_2 f_2,\ldots,\e_n f_n)$, in its domain of definition
962: $C\subset {\cal R}^n$.}
963:
964: \*
965: \0{\it Remark:} Note that the {\it r.h.s.} of Eq.(\ref{7.4}) {\it does not
966: depend on $f_2,\ldots,f_n$}. The independence has been exploited in
967: \cite{Ga96a} to show that when the forcing on the system is due to several
968: forces of respective intensities $E_1,\ldots,E_s$ then by taking $F_1=
969: \frac{\s(x)}{\s_+},\, F_2=\partial_{E_k}\s(x)$, the Eq.(\ref{7.4}) implies,
970: setting $j_k(x)=\partial_{E_k}\s(x)$ and $J_k=\media{j_k}_\m$, the Green
971: Kubo relations (hence Onsager reciprocity)
972:
973: \be L_{hk}=\partial_{E_h} J_k\big|_{\V E=\V0}=\frac12\int_{-\infty}^\infty
974: \media{j_h(S_\t
975: x)\,j_k(x)}_{E=0}dt=L_{kh}.\label{7.5}\ee
976: %
977: Therefore FT can be regarded as an {\it extension} to a nonlinear
978: regime of Onsager reciprocity and of the Fluctuation-Dissipation
979: theorems. Such a relation was pointed out in the context of volume
980: preserving dynamics (hence in absence of dissipation), see comments in
981: \cite[p.452]{BK81a} in particular. But it is not clear how to obtain
982: from \cite{BK81a} the dissipative case results in
983: Eq.(\ref{7.1}),(\ref{7.4}),(\ref{7.5}) without the CH. \*
984:
985: \def\SEC{Patterns and Onsager-Machlup Fluctuations}
986: \section{Fluctuation Patterns, Onsager-Machlup Theory}
987: \label{sec8}\iniz
988:
989: The last comment makes it natural to inquire whether there are more
990: direct and physical interpretations of the FT (hence of the
991: meaning of CH) when the external forcing is really different from the
992: value $0$ (the value always assumed in Onsager's theory).
993:
994: The proof of the FT allows, as well, to deduce, \cite{Ga99}, an
995: apparently more general statement (closely related to a relation
996: recently found in the theory of the Kraichnan model of $2$-dimensional
997: turbulence and called ``multiplicative'' fluctuation theorem,
998: \cite{CDG07}) which can be regarded as an extension to nonequilibrium
999: of the Onsager-Machlup theory of fluctuation patterns.
1000:
1001: Consider observables $\V F=(F_1\defi {\s}/{\s_+},\ldots,F_n)$ which
1002: have a well defined time reversal parity: $F_i(Ix)=\e_{F_i} F_i(x)$,
1003: with $\e_{F_i}=\pm1$. Let $F_{i+}$ be their time average ({\it i.e.}
1004: their SRB average) and let $t\to \Bf(t)=(\f_1(t),\ldots,$ $\f_n(t))$
1005: be a smooth bounded function. Look at the probability, relative to
1006: the SRB distribution ({\it i.e.} in the ``natural stationary state'')
1007: that $F_i(S_t x)$ is $\f_i(t)$ for $t\in [-\frac\t2,\frac\t2]$: we say
1008: that $\V F$ ``follows the fluctuation pattern'' $\Bf$ in the time
1009: interval $t\in [-\frac\t2,\frac\t2]$.
1010:
1011: No assumption on the fluctuation size, nor on the size of the forces
1012: keeping the system out of equilibrium, will be made. Besides the
1013: CH we assume, however, that the evolution is time
1014: reversible {\it also} out of equilibrium and that the phase space
1015: contraction rate $\s_+$ is not zero (the results hold no
1016: matter how small $\s_+$ is and, appropriately interpreted, they make
1017: sense even if $\s_+=0$, but in that case they become trivial).
1018:
1019: We denote $\z(p,\Bf)$ the {\it large deviation function} for observing
1020: in the time interval $[-\frac\t2,\frac\t2]$ an average phase space
1021: contraction $\s_\t\defi\frac1\t\int_{-\t/2}^{\t/2}\s(S_tx)dt= p\s_+$
1022: and at the same time a fluctuation pattern $\V F(S_tx)=\Bf(t)$. This
1023: means that the probability that the {\it dimensionless phase space
1024: contraction rate} $p$ is in a closed set $\D$ and $F$ is in a closed
1025: neighborhood of an assigned $\Bps$,\footnote{\nota By ``closed
1026: neighborhood'' $U_{\Bps,\e}$, $\e>0$, around $\Bps$, we mean that
1027: $|F_i(S_tx)-\ps_i(t)|\le\e$ for $t\in[-\frac\t2,\frac\t2]$.} denoted
1028: $U_{\Bps,\,\e}$, is given by:
1029:
1030: \be \exp\Big(\sup_{p\in\D,\Bf\in U_{\Bps,\e}}
1031: {\t\,\z(p,\Bf)}\Big)\label{8.1}\ee
1032: %
1033: to leading order as $\t\to\infty$ ({\it i.e.} the logarithm of the mentioned
1034: probability divided by $\tau$ converges as $\t\to\infty$ to
1035: $\sup_{p\in\D,\Bf\in U_{\Bps,\e}} \z(p,\Bf)$). Needless to say $p$ and
1036: $\Bf$ have to be ``possible'' otherwise $\z$ has to be set $-\infty$, as
1037: in the FT case in Sec.\ref{sec6}, comment (6).
1038:
1039: Given a reversible, dissipative, transitive Anosov flow the
1040: fluctuation pattern $t\to\Bf(t)$ and the time reversed pattern
1041: $t\to\e_F\Bf(-t)$ are then related by the following:
1042: \*
1043:
1044: \0{\bf Conditional reversibility relation:} {\sl If $\V
1045: F=(F_1,\ldots,F_n)$ are $n$ observables with defined time reversal
1046: parity $\e_{F_i}=\pm1$ and if $\t$ is large the fluctuation pattern
1047: $\Bf(t)$ and its time reversal $I\f_i(t)\defi \e_{F_i}\f_i(-t)$ will
1048: be followed with equal likelihood if the first is conditioned to a
1049: contraction rate $p$ and the second to the opposite $-p$. This
1050: holds because:
1051:
1052: \be \frac{\z(p,\Bf)-\z(-p,I\Bf) }{p\s_+}=1 \qquad {\rm for\ } |p|\le
1053: p^*\label{8.2}\ee
1054: %
1055: with $\z$ introduced in Eq.(\ref{8.1}) and a suitable $p^*\ge1$.}
1056: \*
1057:
1058: It will appear, in Sec.\ref{sec9}, that the phase space contraction
1059: rate should be identified with a macroscopic quantity,
1060: the {\it entropy creation rate}. Then
1061: the last theorem can be interpreted as saying, in other words, that
1062: while it is very difficult, in the considered systems, to see an
1063: ``anomalous'' average entropy creation rate during a time $\t$ ({\it
1064: e.g.} $p=-1$), it is also true that ``that is the hardest thing to
1065: see''. Once we see it, {\it all the observables will behave
1066: strangely} and the relative probabilities of time reversed patterns
1067: will become as likely as those of the corresponding direct patterns
1068: under ``normal'' average entropy creation regime.
1069:
1070: ``A waterfall will go up, as likely as we expect to see it going down,
1071: in a world in which for some reason the entropy creation rate has
1072: changed sign during a long enough time.'' We can also say that the
1073: motion on an attractor is reversible, even in presence of dissipation,
1074: once the dissipation is fixed.
1075:
1076: The result in Eq.(\ref{8.2}) is a ``{\it relation}'' rather than a
1077: theorem because, even in the hyperbolic cases, the precise
1078: restrictions on the ``allowed'' test functions $\f_i(t)$ have not been
1079: discussed in \cite{Ga99} from a strict mathematical viewpoint and it
1080: would be interesting to formulate them explicitly and investigate
1081: their generality.\footnote{\nota A sufficient condition should be that
1082: $\f_i(t)$ are bounded and smooth.}\*
1083:
1084: The result can be informally stated in a only apparently
1085: stronger form, for $|p|<p^*$, and with the warnings in remark (4)
1086: preceding the analogous Eq.(\ref{5.3}), as
1087:
1088: \be \frac{P_\t(\hbox{\rm
1089: for all}\ j, {\rm and}\, t\in[-\frac12\t,\frac12\t] \,:
1090: F_j(S_tx)\sim\f_j(t))} {P_\t(\hbox{\rm for all}\ \,j, {\rm and}\,
1091: t\in[-\frac12\t,\frac12\t]\,: F_j(S_tx)\sim -\f_j(-t))}
1092: =e^{\,p\,\s_+\, \t+O(1)},\label{8.3}\ee
1093: %
1094: where $P_\t$ is the SRB probability, {\it provided the phase space
1095: contraction $\s(x)$ is a function of the observables $\V F$}. This is
1096: certainly the case if $\s$ is one of the $F_i$, for instance if
1097: $\s=F_1$. Here $F_j(S_tx)\sim \f_j(t)$ means
1098: $|F_j(S_tx)-\f_j(t)|$ small for $t\in[-\frac\t2,\frac\t2]$.
1099: \*
1100:
1101: \0{\it Remarks:}
1102: %
1103: \\
1104: %
1105: (1) A relation of this type has been remarked
1106: recently in the context of the theory of Lagrangian trajectories in
1107: the Kraichnan flow, \cite{CDG07}.
1108: %
1109: \\
1110: %
1111: (2) One should note that in applications results like Eq.(\ref{8.3})
1112: will be used under the CH and therefore other errors may arise because
1113: of its approximate validity (the hypothesis in fact essentially states
1114: that ``things go as if'' the system was hyperbolic): they may depend
1115: on the number $N$ of degrees of freedom and we do not control them
1116: except for the fact that, if present, their relative value should tend
1117: to $0$ as $N\to\infty$: there may be (and there are) cases in which
1118: the chaotic hypotesis is not reasonable for small $N$ ({\it e.g.}
1119: systems like the Fermi-Pasta-Ulam chains) but it might be correct for
1120: large $N$. We also mention that, on the other hand, for some systems
1121: with small $N$ the CH may be already regarded as valid ({\it e.g.} for
1122: the models in \cite{CELS93}, \cite{ECM93,BGG97}).
1123: %
1124: \\
1125: %
1126: (3) The proofs of FT and the corollaries are not difficult. Once their
1127: meaning in terms of coarse graining is understood, the a priori rather
1128: misterious SRB distribution $\m$ is represented, surprisingly, as a
1129: Gibbs distribution for a $1$--dimensional spin system, which is
1130: elementary and well understood. In Appendix A1,A2 some details are given
1131: about the nature of coarse graining and in Appendix A3 the steps of
1132: the proof of FT are illustrated.
1133: \*
1134:
1135: In conclusion the FT is a general parameterless relation valid, in
1136: time reversible systems, independently of the number of degrees of
1137: freedom: the CH allows us to consider it as a manifestation of time
1138: reversal symmetry.
1139:
1140: \def\SEC{Reversible thermostats and Entropy Creation}
1141: \section{Reversible thermostats and Entropy Creation}
1142: \label{sec9}\iniz
1143:
1144: Recalling that kinetic theory developed soon after the time average of
1145: a mechanical quantity, namely kinetic energy, was understood to
1146: have the meaning of absolute temperature, it is tempting to consider
1147: quite important that, from the last three decades of research on
1148: nonequilibrium statistical Mechanics, an interpretation emerged
1149: of the physical meaning of the mechanical quantity $\s$ = phase space
1150: contraction.
1151:
1152: A system in contact with thermostats can generate entropy in the sense
1153: that it can send amounts of heat into the thermostats thus increasing
1154: their entropy by the ratio of the heat to the temperature, because the
1155: thermostats must be considered in thermal equilibrium.
1156:
1157: Furthermore if phase space contraction can be identified with a
1158: physical quantity, accessible by means of calorimetric/thermometric
1159: measurements, then the FT prediction becomes relevant and observable
1160: and the CH can be subjected to tests, {\it independently on the
1161: microscopic model that one may decide to assume}, which therefore
1162: become possible also in real experiments.
1163:
1164: It turns out that in very general thermostat models entropy production
1165: rate can be identified with phase space contraction {\it up to a
1166: ``total time derivative''}: and since additive total time derivatives
1167: (as we shall see) do not affect the asympotic distribution of time
1168: averages, one can derive a FR for the entropy production (a quantity
1169: accessible to measurement) from a FR for phase space contraction (a
1170: quantity, in general, {\it not accessible} except in numerical
1171: simulations, because it requires a precise model for the system, as a
1172: rule not available).
1173:
1174: As an example, of rather general nature, consider the following
1175: one, obtained by imagining a system which is in contact with thermostats
1176: that are ``external'' to it. The particles of the system $\CC_0$ are
1177: enclosed in a container, also called $\CC_0$, with elastic boundary
1178: conditions surrounded by a few thermostats which consist of particles,
1179: all of unit mass for simplicity, interacting with the system via short
1180: range interactions, through a portion $\partial_i{\CC_0}$ of the surface
1181: of ${\CC_0}$, and subject to the constraint that the total kinetic
1182: energy of the $N_i$ particles in the $i$-th thermostat is
1183: $K_i=\frac{1}2 \dot{\V X}_i^2=\frac32 N_i k_B T_i$. A symbolic
1184: illustration is in Fig.1.
1185:
1186:
1187: %\input fig
1188: \eqfig{110}{90}{}{fig.eps}{}
1189:
1190: \0{\nota Fig.1: Particles in $\CC_0$ (``system particle'') interact
1191: with the particles in the shaded regions (``thermostats particles'');
1192: the latter are constrained to have a fixed total kinetic energy.} \*
1193:
1194: The equations of motion will be (all masses equal for simplicity)
1195:
1196: \be \eqalign{
1197: m\ddot{\V X}_0=&-\partial_{\V X_0}\Big( U_0(\V X_0)+\sum_{j>0}
1198: W_{0,j}(\V X_{0},\V X_j)\Big)+\V E(\V X_0),
1199: \cr
1200: m\ddot{\V X}_i=&-\partial_{\V X_i}\Big( U_i(\V X_i)+
1201: W_{0,i}(\V X_{0},\V X_i)\Big)-\a_i \dot{\V X}_i\hbox{\hglue1.1truecm}
1202: \cr}\label{9.1}\ee
1203: %
1204: with $\a_i$ such that $K_i$ is a constant. Here $W_{0,i}$ is the
1205: interaction potential between particles in $\CC_i$ and in $\CC_0$,
1206: while $U_0,U_i$ are the internal energies of the particles in
1207: $\CC_0,\CC_i$ respectively. We imagine that the energies $W_{0,j},U_j$
1208: are due to {\it smooth} translation invariant pair
1209: potentials; repulsion from the boundaries of the containers will be
1210: elastic reflection.
1211:
1212: It is assumed, in Eq.(\ref{9.1}), that there is no direct interaction
1213: between different thermostats: their particles interact directly only
1214: with the ones in $\CC_0$. Here $\V E({\V X}_0)$ denotes possibly
1215: present external positional forces stirring the particles in
1216: $\CC_0$. The contraints on the thermostats kinetic energies give
1217:
1218: \be \a_i\equiv \frac{Q_i-\dot U_i}{3N_i k_B T_i}\qquad \otto\qquad
1219: K_i\equiv const\defi\frac32 N_i k_B T_i\label{9.2}\ee
1220: %
1221: where $Q_i$ is the work per unit time that particles outside the thermostat
1222: $\CC_i$ (hence in $\CC_0$) exercise on the particles in it, namely
1223:
1224: \be Q_i\defi-\dot{\V
1225: X}_i\cdot\partial_{{\V X}_{i}}W_{0,i}(\V X_{0},{\V X}_i)\label{9.3}\ee
1226: %
1227: and it will be interpreted as the ``{\it amount of heat}'' $Q_i$ entering
1228: the thermostat $\CC_i$ per unit time.
1229:
1230: The main feature of the model is that the thermostats are
1231: external to the system proper: this makes the model suitable for the
1232: study of situations in which no dissipation occurs in the interior of
1233: a system but it occurs only on the boundary.
1234:
1235: The {\it divergence} $-\s(\dot{\V X},\V X)$ of the equations of
1236: motion, which gives the rate of contraction of volume elements around
1237: $d\dot{\V X}d\V X$, does not vanish and can be computed in the model
1238: in Fig.1; simple algebra yields, remarkably,
1239:
1240: \be \eqalign{
1241: \s(\V{{\dot X}},\V X)=&\,\e(\V{{\dot X}},\V X)+\dot R(\V X),\cr
1242: \e(\V{{\dot X}}, \V X)=&\sum_{j>0} \frac{Q_j}{k_B T_j},
1243: \qquad R(\V X)= \sum_{j>0} \frac{U_j}{k_B T_j}\cr}
1244: \label{9.4}\ee
1245: %
1246: where $\e(\V{{\dot X}},\V X)$ can be interpreted as the {\it entropy
1247: production rate}, because of the meaning of $Q_i$ in Eq.(\ref{9.3}).%
1248: \footnote{\nota Eq.(\ref{9.4}) are correct up to $O(N^{-1})$ if
1249: $N=\min N_j$ because the addends should contain also a factor
1250: $(1-\frac1{3 N_j})$ to be exact: for simplicity $O(1/N)$ corrections
1251: will be ignored here and in he following (their inclusion would imply
1252: trivial changes without affecting the physical interpretation),
1253: \cite{Ga06c}.\label{7}\label{footnote7}}%for page reference O(1/N)
1254:
1255: This is an interesting result because of its generality: it has
1256: implications for the thermostated system considered in Fig.1, for
1257: instance. It is remarkable that the quantity $p$ has a simple physical
1258: interpretation: Eq.(\ref{9.1}) shows that the functions $\z_\s(p)$ and
1259: $\z_\e(p)$ are {\it identical} because, since $R$ is bounded by our
1260: assumption of smoothness, Eqs.
1261: (\ref{9.2}) and (\ref{9.3}) imply
1262:
1263: \be \frac1\t \int_0^\t {\s(S_t(\dot{\V X},\V X))}dt
1264: \equiv \frac1\t \int_0^\t
1265: {\e(S_t(\dot{\V X},\V X))}dt +\frac{R(\t)-R(0)}{\t},\label{9.5}\ee
1266: %
1267: so that
1268:
1269: \be \s_+=\lim_{\t\to\infty}\frac1\t \int_0^\t \s(S_t(\dot{\V X},\V X))dt\equiv
1270: \lim_{\t\to\infty}\frac1\t \int_0^\t \e(S_t(\dot{\V X},\V X))dt=\e_+
1271: \label{9.6}\ee
1272: %
1273: and the asmptotic distributions of
1274:
1275: \be p'=\frac1\t \int_0^\t
1276: \frac{\s(S_t(\dot{\V X},\V X))}{\s_+}dt, \qquad\hbox{and of}\qquad
1277: p=\frac1\t \int_0^\t
1278: \frac{\e(S_t(\dot{\V X},\V X))}{\e_+}dt\label{9.7}\ee
1279: %
1280: {\it are the same}.
1281: %
1282:
1283: The Eq.(\ref{9.1}) are time reversible (with $I(\dot{\V X},\V
1284: X)=(-\dot{\V X},\V X)$): then under the CH the large deviations rate
1285: $\z(p)$ for the observable $\frac{\s}{\s_+}$ satisfies the ``{\it
1286: fluctuation relation}'', Eq.(\ref{7.1}). It also follows that the
1287: large deviations rate for $\frac{\e}{\e_+}$, identical to $\z(p)$,
1288: satisfies it as well.
1289:
1290: The point is that $\e$ is measurable by ``calorimetric and
1291: thermometric measurements'', given its interpretation of entropy
1292: increase of the thermostats. Therefore the CH can be subjected to
1293: test or it can be used to ``predict'' the frequency of occurence of
1294: unlikely fluctuations.
1295: \*
1296:
1297: \0{\it Comment:} This is a rather general example of thermostats
1298: action, but it is just an example. For instance it can be generalized
1299: further by imagining that the system is thermostatted in its
1300: interior. A situation that arises naturally in the theory of electric
1301: conduction. In the latter case the electrons move across the lattice
1302: of the metal atoms and the lattice oscillations, {\it i.e.} the
1303: phonons, absorb or give energy. This can be modeled by adding a
1304: ``inner'' thermostat force $-\a_0\dot{\V x}_i$, acting on the
1305: particles in $\CC_0$, which fixes the temperature of the electron
1306: gas. Actualy a very similar model appeared in the early days of
1307: Statistical Mechanics, in Drude's theory of electric conductivity,
1308: \cite{Be64}. Other examples can be found in \cite{Ga06c}.
1309:
1310: \def\SEC{Fluids}
1311: \section{Fluids}
1312: \label{sec10}\iniz
1313:
1314: The attempt to put fluids and turbulence within the context provided
1315: by the ideas exposed in the previous sections forces to consider cases
1316: in which dissipation takes place irreversibly. This leads us to a few
1317: conjectures and remarks.
1318:
1319: To bypass the obstacle due to the nonreversibility of the fluid
1320: equations which, therefore, seem quite far from the equations
1321: controlling the thermostated systems just considered, the
1322: following ``equivalence conjecture'', \cite{Ga02}, has been
1323: formulated. Consider the two equations for an incompressible flow
1324: with velocity field $\V u(\V x,t)$, $\BDpr\cdot\V u=0$, in
1325: periodic boundary condition for simplicity,
1326:
1327: \be \eqalign{
1328: &\dot{\V u}+\T {\V u}\cdot \T\partial\, {\V u}=\n \D\V u-\partial p+ \V
1329: g,\cr
1330: &\dot{\V u}+\T {\V u}\cdot \T\partial {\V u}=\a(\V u) \D\V u-\partial p+ \V
1331: g,\cr}\label{10.1}\ee
1332: %
1333: where $\a(\V u)=\frac{\int \V u\cdot \V g\,d\V x}{\int (\partial \V u)^2
1334: \,d\V x}$ is a ``Lagrange multiplier'' determined so that the total
1335: energy $\EE\defi \int \V u^2\,d\V x$ is exactly constant.
1336:
1337: Note that velocity reversal $I:\,\V u(\V x)\to -\V u(\V x)$
1338: anticommutes, in the sense of Eq. (\ref{2.1}), with the time evolution
1339: generated by the second equation (because $\a(I\V u)=-\a(\V u)$),
1340: which means that ``fluid elements'' retrace their paths with opposite
1341: velocity.
1342:
1343: Introduce the ``local observables'' $F(\V u)$ as functions depending only
1344: upon finitely many Fourier components of $\V u$, {\it i.e.} on the ``large
1345: scale'' properties of the velocity field $\V u$.
1346: %
1347: Then, {\it conjecture}, \cite{Ga97b}, the two equations should
1348: have ``same large scale statistics'' in the limit $R\to+\infty$. If
1349: $\m_\n$ and $\wt\m_\EE$ denote the respective SRB distributions of the
1350: first and the second equations in Eq. (\ref{10.2}), by
1351: {\it ``same statistics''} as $R\to\infty$ it is meant that
1352: \*
1353:
1354: \0(1) if the total energy $\EE$ of the initial datum $\V u(0)$ for the
1355: second equation is chosen equal to the average $\media{\int \V
1356: u^2\,d\V x}_{\m_\n}$ for the SRB distribution $\m_\n$ of the first equation,
1357: then
1358: %
1359: \\
1360: %
1361: (2) the two SRB distributions $\m_\n$ and $\wt\m_\EE$ are such
1362: that, in the limit $R\to\infty$, the difference
1363: $\media{F}_{\m_\n}-\media{F}_{{\wt\m}_\EE}\tende{R\to+\infty}0$.
1364:
1365: \*
1366: So far {\it only numerical tests} of the conjecture, in strongly cut off
1367: $2$-dimensional equations, have been attempted (\cite{RS99}).
1368: \*
1369:
1370: An analogy with the termodynamic limit appears naturally: namely the
1371: Reynolds number plays the role of the volume, locality of observables
1372: becomes locality in $\V k$-space, and $\n,\EE$ play the role of
1373: canonical temperature and microcanonical energy of the SRB
1374: distributions of the two different equations in (\ref{10.1}),
1375: respectively $\m_\n$ and $\wt\m_\EE$. \*
1376:
1377: The analogy suggests to question whether reversibility of the second
1378: equation in Eq.(\ref{10.1}) can be detected. In fact to be able to see
1379: for a large time a viscosity opposite to the value $\n$ would be very
1380: unphysical and would be against the spirit of the conjecture.
1381:
1382: If the CH is supposed to hold it is possible to use the FT, which is a
1383: consequence of reversibility, to estimate the probability that, say,
1384: the value of $\a$ equals $-\n$. For this purpose we have to first
1385: determine the attracting set.
1386:
1387: Assuming the K41, \cite{Ga02}, theory of turbulence the attracting set will
1388: be taken to be the set of fields with Fourier components $\V
1389: u_{\V k}=0$ unless $|\V k|\le R^{\frac34}$.
1390:
1391: Then the expected identity $\media{\a}=\n$, between the average
1392: friction $\media\a$ in the second of Eq.(\ref{10.1}) and the viscosity
1393: $\n$ in the first, implies that the divergence of the evolution in the
1394: second of Eq.(\ref{10.1}) is in average
1395:
1396: \be
1397: \s\sim \n \,\sum_{|\V k|\le R^{3/4}}2|\V k|^2\sim\,\n\,(\frac{2\p}L)^2
1398: \frac{8\p}5 R^{15/4}\label{10.2}\ee
1399: \*
1400: By FT the SRB-probability to see, in motions following the second
1401: equation in Eq. (\ref{10.2}), a ``{\it wrong}'' average friction $-\n$
1402: for a time $\t$ is
1403:
1404: \be {\rm Prob}_{srb}\sim \exp{\big(-\t \n \frac{32\p^3}{5L^2} R^{\frac{15}{4}
1405: }\big)}\,\defi\, e^{-g\t}\label{10.3}\ee
1406: %
1407: It can be estimated in the situation considered below for a flow in air:
1408:
1409: \be \left\{\eqalign{
1410: \n=& 1.5\,10^{-2}\,\frac{cm^2}{sec},\quad v=10.\,\frac{cm}{sec}\,
1411: \quad L=100.\,cm\cr
1412: R=& 6.67\,10^{4},\quad g=3.66\,10^{14}\, sec^{-1}\cr
1413: P\defi& {\rm Prob}_{srb}=
1414: e^{-g\t}=e^{-3.66\,10^8},\qquad {\rm if}\quad
1415: \t=10^{-6}\cr}\right.\label{10.4}\ee
1416: %
1417: where the first line are data of an example of fluid motion and the
1418: other two lines follow from Eq.(\ref{10.3}). They show that, by FT,
1419: viscosity can be $-\n$ during $10^{-6}s$ ({\it say}) with probability
1420: $P$ as in Eq.(\ref{10.4}): unlikelyhood is similar in spirit to the
1421: estimates about Poincar\'e's recurrences, \cite{Ga02}. \*
1422:
1423: \0(2) If we imagine that the particles are so many that the system can
1424: be well described by a macroscopic equation, like for instance the NS
1425: equation, then there will be two ways of computing the entropy
1426: creation rate. The first would be the classic one described for
1427: instance in \cite{DGM84}, and the second would simply be the
1428: divergence of the microscopic equations of motion in the model of
1429: Fig.1, under the assumption that the motion is closely described by
1430: macroscopic equations for a fluid in local thermodynamic equilibrium,
1431: like the NS equations. This can be correct in the limit in which space
1432: and time are rescaled by $\e$ and $\e^2$ and the velocity field by
1433: $\e$, and $\e$ is small. Since local equilibrium is supposed, it
1434: will make sense to define a local entropy density $s(\V x)$ and a
1435: total entropy of the fluid $S=\int s(\V x)\,d\V x$.
1436:
1437: The evaluation of the expression for the entropy creation rate as a
1438: divergence $\s$ of the microscopic equations of motion leads to,
1439: \cite{Ga06}, a value $\media\e$ with average (over a microscopically
1440: long time short with respect to the time scale of the fluid evolution)
1441: related to the classical entropy creation rate in a NS fluid as
1442:
1443: \be \eqalign{k_B \media\e=&k_B\e_{classic}+\dot S, \cr
1444: k_B \e_{classic}=&\int_{\CC_0}\Big(\k\,
1445: \big(\frac{\V\partial T}{T}\big)^2
1446: +\h\, \frac1T{\W{\Bt}'\cdot\W\partial \V u}\Big)\,d\V x\cr}\label{10.5}\ee
1447: %
1448: where $\W{\Bt}'$ is the tensor $(\partial_i u_j+\partial_j u_i)$ and
1449: $\h$ is the dynamic viscosity, so that the two expressions differ by
1450: the time derivative of an observable, which equals the total
1451: equilibrium entropy of the fluid $S=\int s(\V x)\, d\V x$ where $s $
1452: is the thermodynamical entropy density in the assumption of local
1453: equilibrium; see comment on additive total derivatives preceding
1454: Fig.1.
1455:
1456:
1457: \def\SEC{Quantum Systems}
1458: \section{Quantum Systems}
1459: \label{sec11}\iniz
1460:
1461: Recent experiments deal with properties on mesoscopic and atomic
1462: scale. In such cases the quantum nature of the systems may not always
1463: be neglected, paricularly at low temperature, and the question is
1464: whether a fluctuation analysis parallel to the one just seen in the
1465: classical case can be performed in studying quantum phenomena.
1466:
1467: Thermostats have, usually, a macroscopic phenomenological nature: in a
1468: way they should be regarded as classical macroscopic objects in which
1469: no quantum phenomena occur. Therefore it seems natural to model them
1470: as such and define their temperature as the average kinetic
1471: energy of their constituent particles so that the question of how to define
1472: it does not arise.
1473:
1474: Consider the system in Fig.1 when the quantum nature of the particles
1475: in $\CC_0$ cannot be neglected. Suppose for simplicity (see \cite{Ga07})
1476: that the nonconservative force $\V E(\V X_0)$ acting on $\CC_0$
1477: vanishes, {\it i.e.} consider the problem of heat flow through
1478: $\CC_0$. Let $H$ be the operator on $L_2(\CC_0^{3N_0})$, space of
1479: symmetric or antisymmetric wave functions $\Ps(\V X_0)$,
1480:
1481: \be H=
1482: -\frac{\hbar^2}{2m}\D_{\V X_0}+ U_0(\V X_0)+\sum_{j>0}\big(U_{0j}(\V X_0,\V
1483: X_j)+U_j(\V X_j)+K_j\big)\label{11.1}\ee
1484: %
1485: where $\D_{\V X_0}$ is the Laplacian, and note that its spectrum
1486: consists of eigenvalues $E_n=E_n(\{\V X_j\}_{j>0})$, for $\V X_j$
1487: fixed (because the system in $\CC_0$ has finite size).
1488:
1489: A system--reservoirs model can be the {\it dynamical system} on the
1490: space of the variables $\big(\Ps,(\{\V X_j\},$ $\{\V{{\dot
1491: X}}_j\})_{j>0}\big)$ defined by the equations (where
1492: $\media{\cdot}_\Ps\,=$ expectation in the state $\Ps$)
1493:
1494: \be \eqalign{
1495: -i\hbar {\dot\Ps(\V X_0)}=& \,(H\Ps)(\V X_0),\kern20mm{\rm and\ for}\
1496: j>0\cr
1497: \V{{\ddot X}}_j=&-\Big(\partial_j U_j(\V X_j)+
1498: \media{\partial_j U_j(\V X_0,\V X_j)}_\Ps\Big)-\a_j \V{{\dot X}}_j\cr
1499: \a_j\defi&\frac{\media{W_j}_\Ps-\dot U_j}{2 K_j}, \qquad
1500: W_j\defi -\V{{\dot X}}_j\cdot \V\partial_j U_{0j}(\V X_0,\V
1501: X_j)}\label{11.2} \ee
1502: %
1503: here the first equation is Schr\"odinger's equation, the second is an
1504: equation of motion for the thermostats particles similar to the one in
1505: Fig.1, (whose notation for the particles labels is adopted here
1506: too). The model has no pretention of providing a physically correct
1507: representation of the motions in the thermostats nor of the
1508: interaction system thermostats, see comments at the end of this
1509: section.
1510:
1511: Evolution maintains the thermostats kinetic energies $K_j\equiv
1512: \frac12\V{{\dot X}}_j^2$ exactly constant, so that they will be used
1513: to define the thermostats temperatures $T_j$ via $K_j=\frac32 k_B T_j
1514: N_j$, as in the classical case.
1515:
1516: Let $\m_0(\{d\Ps\})$ be the {\it formal} measure on
1517: $L_2(\CC_0^{3N_0})$
1518:
1519: \be \Big(\prod_{\V X_0} d\Ps_r(\V X_0)\,d\Ps_i(\V X_0)
1520: \Big)\,\d\Big(\int_{\CC_0} |\Ps(\V Y)|^2\, d\V Y-1\Big)
1521: \label{11.3}\ee
1522: %
1523: with $\Ps_r,\Ps_i$ real and imaginary parts of $\Ps$. The meaning of
1524: (\ref{11.3}) can be understood by imagining to introduce an
1525: orthonormal basis in the Hilbert space and to ``cut it off'' by
1526: retaining a large but finite number $M$ of its elements, thus turning
1527: the space into a high dimensional space $C^M$ (with $2M$ real
1528: dimensions) in which $d\Ps=d\Ps_r(\V X_0)\,d\Ps_i(\V X_0)$ is simply
1529: interpreted as the normalized euclidean volume in
1530: $C^M$.
1531:
1532: The formal phase space volume element $\m_0(\{d\Ps\})\times\n(d\V
1533: X\,d\V{{\dot X}})$ with
1534:
1535: \be \n(d\V X\,d\V{{\dot X}})\defi\prod_{j>0} \Big(\d(\V{{\dot
1536: X}}^2_j-3N_jk_B T_j)\,d\V X_j\,d\V{{\dot X}}_j\Big)
1537: \label{11.4}\ee
1538: %
1539: is conserved, by the unitary property of the wave
1540: functions evolution, just as in the classical case, {\it up
1541: to the volume contraction in the thermostats}, \cite{Ga06c}.
1542:
1543: If $Q_j\defi\media{W_j}_\Ps$ and $R$ is as in Eq.(\ref{9.4}), then the
1544: contraction rate $\s$ of the volume element in Eq.(\ref{11.4}) can be
1545: computed and is (again) given by Eq.(\ref{9.4}) with $\e$, that will
1546: be called {\it entropy production rate}:
1547: setting $R(\V X)\defi \sum_{j>0} \frac{U_j(\V X_j)}{k_B T_j}$, it is
1548:
1549: \be\s(\Ps,\V{{\dot X}},\V X)=\,\e(\Ps,\V{{\dot X}},\V X)+\dot R(\V X),\qquad
1550: \e(\Ps,\V{{\dot X}}, \V X)=\sum_{j>0} \frac{Q_j}{k_B T_j},
1551: \label{11.5}\ee
1552:
1553: In general solutions of Eq.(\ref{11.2}) {\it will not be quasi periodic} and
1554: the Chaotic Hypothesis, \cite{GC95b,Ga00,Ga07}, can be assumed: if so the
1555: dynamics should select an SRB distribution $\m$. The
1556: distribution $\m$ will give the statistical properties of the
1557: stationary states reached starting the motion in a thermostat
1558: configuration $(\V X_j,\V{{\dot X}}_j)_{j>0}$, randomly chosen with
1559: ``uniform distribution'' $\n$ on the spheres $m\V{{\dot X}}_j^2=3N_jk_B
1560: T_j$ and in a random eigenstate of $H$. The distribution $\m$, if
1561: existing and unique, could be named the {\it SRB distribution}
1562: corresponding to the chaotic motions of Eq.(\ref{11.2}).
1563:
1564: In the case of a system {\it interacting with a single thermostat} at
1565: temperature $T_1$ the latter distribution should be equivalent to the
1566: canonical distribution, up to boundary terms.
1567:
1568: Hence an important consistency check, for proposing Eq.(\ref{11.2}) as
1569: a model of a thermostated quantum system, is that there should exist
1570: at least one stationary distribution equivalent to the canonical
1571: distribution at the appropriate temperature $T_1$ associated with the
1572: (constant) kinetic energy of the thermostat: $K_1=\frac32 k_B
1573: T_1\,N_1$. In the corresponding classical case this is an established
1574: result, \cite{EM90,Ga00,Ga06c}.
1575:
1576: A natural candidate for a stationary distribution could be to
1577: attribute a probability proportional to $d\Ps\,d\V X_1\,d \dot{\V
1578: X}_1$ times
1579:
1580: \be
1581: \sum_{n=1}^\infty e^{-\b_1 E_n}\d(\Ps-\Ps_n(\V
1582: X_1)\,e^{i\f_n})\,{d\f_n}\,\d(\dot{\V X}_1^2-2K_1)\label{11.6}\ee
1583: %
1584: where $\b_1=1/k_B T_1$, $\Ps$ are wave functions for the system in
1585: $\CC_0$, ${\dot {\V X}_1, \V X_1}$ are positions and velocities of
1586: the thermostat particles and $\f_n\in [0,2\p]$ is a phase, $E_n=E_n(\V
1587: X_1)$ is the $n$-th level of $H(\V X_1)$, with $\Ps_n(\V X_1)$ the
1588: corresponding eigenfunction. The average value of an observable $O$
1589: for the system in $\CC_0$ in the distribution $\m$ in (\ref{11.6}) would
1590: be
1591:
1592: \be \media{O}_\m=Z^{-1}\int {\rm Tr}\, (e^{-\b H(\V X_1)} O)\,\d(\dot{\V
1593: X}_1^2-2K_1)d\V X_1\,d \dot{\V X}_1\label{11.7}\ee
1594: %
1595: where $Z$ is the integral in (\ref{11.7}) with $1$ replacing $O$,
1596: (normalization factor). Here one recognizes that $\m$ attributes to
1597: observables the average values corresponding to a Gibbs state at
1598: temperature $T_1$ with a random boundary condition $\V X_1$.
1599:
1600: However Eq.(\ref{11.6}) {\it is not invariant} under the evolution
1601: Eq.(\ref{11.2}) and it seems difficult to exhibit explicitly an
1602: invariant distribution. Therefore one can say that the SRB
1603: distribution for the evolution in (\ref{11.2}) is equivalent to the
1604: Gibbs distribution at temperature $T_1$ only as a conjecture.
1605:
1606: Nevertheless it is interesting to remark that under the {\it adiabatic
1607: approximation} the eigenstates of the Hamiltonian at time $0$ evolve
1608: by simply following the variations of the Hamiltonian $H(\V X(t))$ due
1609: to the motion of the thermostats particles, without changing quantum
1610: numbers (rather than evolving following the Schr\"odinger equation and
1611: becoming, therefore, different from the eigenfunctions of $H(\V
1612: X(t))$).
1613:
1614: In the adiabatic limit in which the classical motion of the
1615: thermostat particles takes place on a time scale much slower than the
1616: quantum evolution of the system the distribution (\ref{11.6}) {\it is
1617: invariant}.
1618:
1619: This can be checked by first order perturbation analysis
1620: which shows that, to first order in $t$, the variation of the energy
1621: levels (supposed non degenerate) is compensated by the phase space
1622: contraction in the thermostat, \cite{Ga07}.
1623: %
1624: Under time evolution, $\V X_1$ changes, at time $t>0$, into $\V X_1+t
1625: \V{{\dot X}}_1+O(t^2)$ and, assuming non degeneracy, the eigenvalue
1626: $E_n(\V X_1)$ changes, by perturbation analysis, into $E_n+t \,
1627: e_n+O(t^2)$ with
1628:
1629: \be e_n\defi t\media{\V{{\dot X}}_1\cdot\V\partial_{\V X_1}
1630: U_{01}}_{\Ps_n}+t \V{{\dot X}}_1\cdot\V\partial_{\V X_1}
1631: U_{1}=-t\,(\media{W_1}_{\Ps_n}+\dot R_1)=-\frac1{\b_1}\a_1.\label{11.8}
1632: \ee
1633: %
1634: Hence the Gibbs factor changes by $e^{-\b t e_n}$ and at the same time
1635: phase space contracts by $e^{t \frac{3 N_1 e_n}{2K_1}}$, as it follows
1636: from the expression of the divergence in Eq.(\ref{11.5}). {\it
1637: Therefore if $\b$ is chosen such that $\b=(k_B T_1)^{-1}$ the state
1638: with distribution Eq.(\ref{11.6}) is stationary}, (recall that for
1639: simplicity $O(1/N)$, see footnote${}^{\ref{footnote7}}$ on
1640: p.\pageref{7} %%%% for page ref O(1/N)
1641: is neglected). This shows that, {\it in the adiabatic approximation},
1642: interaction with only one thermostat at temperature $T_1$ admits at
1643: least one stationary state. The latter is, by construction, a Gibbs
1644: state of thermodynamic equilibrium with a special kind (random $\V
1645: X_1,\V{{\dot X}}_1$) of boundary condition and temperature $T_1$. \*
1646:
1647: \0{\it Remarks:}
1648: (1) The interest of the example is to show that even in quantum
1649: systems the chaotic hypothesis makes sense and the intepretation
1650: of the phase space contraction in terms of entropy production
1651: remains unchanged. In general, under the chaotic hypothesis, the
1652: SRB distribution of (\ref{11.2}) (which in presence of forcing, or
1653: of more than one thermostat is certainly quite not trivial, as in
1654: the classical Mechanics cases) will satisfy the fluctuation
1655: relation because the fluctuation theorem only depends on
1656: reversibility: so the model (\ref{11.2}) might be suitable (given
1657: its chaoticity) to simulate the steady states of a quantum system
1658: in contact with thermostats.
1659:
1660:
1661: \0(2) It is certainly unsatisfactory that a stationary distribution
1662: cannot be explicitly exhibited for the single thermostat case (unless
1663: the adiabatic approximation is invoked). However, according to the
1664: proposed extension of the CH, the model does
1665: have a stationary distribution which should be equivalent (in the sense
1666: of ensembles equivalence) to a Gibbs distribution at the same
1667: temperature.
1668:
1669: \0(3) The non quantum nature of the thermostat considered here and the
1670: specific choice of the interaction term between system and
1671: thermostats should not be important: the very notion of thermostat
1672: for a quantum system is not at all well defined and it is natural
1673: to think that in the end a thermostat is realized by interaction
1674: with a reservoir where quantum effects are not
1675: important. Therefore what the analysis really suggests is
1676: that in experiments in which really microscopic systems are
1677: studied the heat exchanges of the system with the external world
1678: should fulfill a FR.
1679:
1680: \0(4) The conjecture can probably be tested with present day
1681: technology. If verified it could be used to develop a
1682: ``Fluctuation Thermometer'' to perform temperature measurements
1683: which are {\it device independent} in the same sense in which the gas
1684: thermometers are device independent ({\it i.e.} do not require
1685: ``calibration'' of a scale and ``comparison''
1686: procedures).
1687: %
1688: \\
1689: %
1690: Consider a system in a stationary state, and imagine inducing small
1691: currents and measuring the average heat output rate $Q_+$ and the
1692: fluctuations in the finite time average heat output rate,
1693: generated by inducing small currents, {\it i.e.} fluctuations of
1694: $p=\frac1\t\int_0^\t \frac{Q(t)}{Q_+}dt$ obtaining the rate function of
1695: $\z(p)$.
1696: %
1697: \\
1698: %
1699: Then it becomes possible to read from the slope of $\z(p)-\z(-p)$,
1700: equal to $\frac{Q_+}{k_B T}$ by the FR, directly the inverse
1701: temperature that the thermostat in contact with the system has:
1702: this could be useful particularly in very small systems (classical
1703: or quantum). The idea is inspired by a similar earlier proposal
1704: for using fluctuation measurements to define temperature in spin
1705: glasses, \cite{CKP97}, \cite[p.216]{CR03}.
1706:
1707:
1708: \section{Experiments ?}
1709: \def\SEC{Experiments ?}
1710: \label{sec12}\iniz
1711:
1712: The (partial) test of the chaotic hypothesis via its implication on
1713: large fluctuations probabilities ({\it i.e.} the fluctuation relation) is
1714: quite difficult. The main reason is that if the forcing is small the
1715: relation degenerates (because $\e_+\to0$) and it can be shown,
1716: \cite{Ga96a}, that to lowest nontrivial order in the size of the forcing it
1717: reduces to the Green-Kubo formula, which is (believed to be) well
1718: established so that the fluctuation relation will not be significant,
1719: being ``true for other reasons'', \cite{DGM84}. See Sec.3.
1720:
1721: Hence one has to consider large forcing. However, under large forcing,
1722: large fluctuations of $p$ become very rare, hence their statistics is
1723: difficult to observe. Furthermore the statistics seems to remain
1724: Gaussian for $p$, in a region around $p=1$ where the data can be
1725: considered reliably unbiased (see below), and until rather large
1726: values of the forcing field or values of $|p-1|$ large compared to the
1727: root mean square deviation $\frac{D}{\sqrt\t}={\langle
1728: (p-1)^2\rangle^{1/2}}$ are reached. Hence
1729: $\z(p)=-\frac1{2D^2} (p-1)^2$ and linearity in $p$ of $\z(p)-\z(-p)$
1730: is trivial. {\it Nevertheless}, in this regime, it follows that
1731: $\frac2{D^2}=\s_+$ which is a nontrivial relation and therefore a
1732: simple test of the fluctuation relation.
1733:
1734: The FR was empirically observed first in such a situation in
1735: \cite{ECM93}, in a simulation, and the first dedicated tests, after
1736: recognizing its link with the CH, were still performed in a Gaussian
1737: regime, so that they were really only tests of $\frac2{D^2}=\s_+$ and of
1738: the Gaussian nature of the observed fluctuations.
1739:
1740: Of course in simulations the forcing can be pushed to ``arbitrarily
1741: large'' values so that the fluctuation relation can, in principle, be
1742: tested in a regime in which $\z(p)$ is sensibly non Gaussian, see
1743: \cite{LLP98}. But far more interesting will be cases in which the
1744: distribution $\z(p)$ is sensibly not Gaussian and which deal with
1745: laboratory experiments rather than simulations. Skepticism towards
1746: the CH is mainly based on the supposed non measurability of the
1747: function $\z(p)$ in the large deviation domain ({\it i.e.} $|p-1|\gg
1748: \sqrt{\langle(p-1)^2\rangle}$).
1749:
1750: In experimental tests several other matters are worrysome,
1751: among which:
1752: \*
1753:
1754: \0(a) is reversibility realized? This is a rather stringent and
1755: difficult point to understand on a case by case basis, because
1756: irreversibility creeps in, inevitably, in dissipative phenomena.
1757:
1758: \0(b) is it allowed to consider $R$, {\it i.e.} the
1759: ``entropy production remainder'' in (\ref{9.3}), bounded? if not there
1760: will be corrections to FR to study (which in some cases,
1761: \cite{CV03a,BGGZ05}, can be studied quite in detail).
1762:
1763: \0(c) does one introduce any bias in the attempts to see statistically
1764: large deviations? for instance in trying to take $\t$ large one may be
1765: forced to look at a restricted class of motions, typically the ones
1766: that remain observable for so long a time. It is easy to imagine that
1767: motions observed by optical means, for instance, will remain within
1768: the field of the camera only for a characteristic time $\t_0$ so that
1769: any statistics on motions that are observed for times $\t>\t_0$ will
1770: be biased (for it would deal with untypical events).
1771:
1772: \0(d) chaotic motions may occur under influence of stochastic
1773: perturbations, so that extensions of FT to stochastic systems may need
1774: to be considered. This is not really a problem because a random
1775: perturbation can be imagined as generated by coupling of the system to
1776: another dynamical system (which, for instance, in simulations would be
1777: the random number generator from which the noise is drawn),
1778: nevertheless it demands careful analysis, \cite{BGG07}.
1779:
1780: \0(e) Nonconvex shape of $\z(p)$, at $|p-1|$ beyond the root mean square
1781: deviation, see Fig.3, is seen often, possibly always, in the
1782: experiments that have been attempted to study large
1783: deviations. Therefore the interpretation of the nonconvexity, via
1784: well understood corrections to FR, seems to be a forced path towards
1785: a full test of the FR, beyond the Gaussian regime, \cite{BGGZ05}.
1786:
1787: \*
1788: \eqfig{0}{125}{\ins{-100}{80}{$10\t_C$}}{fig3.eps}{}
1789:
1790: \0{\nota Fig.3: An histogram of ${\st \log} P_{\t}(p)$, taken from the
1791: data of \cite{BCG07} at time $\t=10\t_C=200$ms: it shows the rather
1792: typical nonconvexity for $|p-\st1|\sim8$ which is of the order of
1793: standard deviation.}
1794: \*
1795:
1796: All the above questions arise in the recent experiment by
1797: Bandi-Cres\-sman-Goldburg, \cite{BCG07}. It encounters all the related
1798: difficulties and to some extent provides the first evidence for the FR
1799: (hence the CH) in a system in which the predictions of the FR are not
1800: the result of a theoretical model which can be solved exactly. The
1801: interpretation of the results is difficult and further investigations
1802: are under way.
1803:
1804: The experiment outcome is not incompatible with FR and, in any event,
1805: it proves that good statistics can be obtained for fluctuations that
1806: extend quite far beyond the root mean square deviation of $p-1$: an
1807: asset of the results in view of more refined experiments.
1808:
1809: A very promising field for experimental tests of the CH and the FR
1810: is granular materials: in granular materials collisions are not
1811: elastic, nevertheless an experiment is proposed in \cite{BGGZ06}.
1812: See comment (6) in Sec.13 and comment (4) to Eq. (\ref{11.8})
1813: for other hints at possible experiments and applications.
1814:
1815:
1816: \section{Comments}
1817: \def\SEC{Comments}
1818: \label{sec13}\iniz
1819:
1820: (1) In the context of the finite thermostats approach, besides systems of
1821: particles subject to deterministic evolution, stochastically evolving
1822: systems can be considered and the FT can be extended to cover the new
1823: situations, \cite{Ku98,LS99,Ma99,CDG07,BGG07}.
1824: %
1825: \\
1826: %
1827: (2) Alternative quantum models have also been considered in the
1828: literature, \cite{Ku00} (stochastic Langevin thermostats), or infinite
1829: thermostats (free and interacting, and possibly with further noise
1830: sources) \cite{FV63,JP02b,HI05,APJP06,JOP07}.
1831: %
1832: \\
1833: %
1834: (3) Many simulations have been performed, starting with the
1835: experiment which showed data that inspired the FT, \cite{ECM93}, and
1836: continuing after the proof of FT and the formulation of the CH, {\it e.g.}
1837: \cite{BGG97}: a few had the purpose of testing the Fr in a nongaussian
1838: regime for the fluctuations of the variable $p$, \cite{LLP98}. In some
1839: cases the results had to be examined closely to understand what was
1840: considered at discrepancy with the FT, \cite{BGGZ05}, (and was not).
1841: %
1842: \\
1843: %
1844: (4) The physical relevance of the particular quantum thermostat model
1845: remains an open question and essentially depends on the conjecture
1846: that the (unknown) SRB distribution for the model in the single
1847: thermostat case is equivalent to the Gibbs distribution at the same
1848: temperature (a property valid in the corresponding classical cases).
1849: Hence the main interest of the model is that it shows that a FR is in
1850: principle possible in finite thermostated quantum systems in
1851: stationary state.
1852: %
1853: \\
1854: %
1855: (5) Few experiments have so far been performed (besides numerical
1856: simulations) to investigate CH and FT: extensions to randomly
1857: forced systems are possible, \cite{Ku98,LS99,Ma99}, and can be
1858: applied to systems that can be studied in laboratory,
1859: \cite{CDGS04,BCG07}: the first experiment designed to test the FR
1860: in a laboratory experiment is the recent work \cite{BCG07}. The
1861: results are consistent with the FR and indicate a promising
1862: direction of research.
1863: %
1864: \\
1865: %
1866: (6) An interesting consequence of the FT is that
1867:
1868: \be \media{e^{-\D S/k_B}}_{srb}\defi
1869: \media{ e^{-\int_0^\t \sum_{j>0}\frac{Q_j(t)}{k_B T_j}} dt}_{srb}= O(1)
1870: \label{13.1}\ee
1871: %
1872: in the sense that the logarithms of both sides divided by $\t$
1873: agree in the limit $\t\to\infty$ ({\it i.e.}
1874: $\lim_{\t\to+\infty}\frac1\t\log \media{e^{\D S/k_B}}=0$) with
1875: corrections of order $O(\frac1\t)$. This has been pointed out by
1876: Bonetto, see \cite{Ga00}, and could have applications in the same
1877: biophysics contexts in which the work theorems, \cite{Ja97,Ja99},
1878: have been applied: for instance one could study stationary heat
1879: exchanges is systems out of equilibrium (rather than measure free
1880: energy differences between equilibrium states at the same
1881: temperature as in \cite{Ja97,Ja99}). The boundedness of the
1882: l.h.s. of Eq. (\ref{13.1}) implied by (\ref{13.1}) can be used to
1883: test whether some heat emissions have gone undetected (which would
1884: imply that the l.h.s. of Eq.(\ref{13.1}) tends to $0$, rather than
1885: staying of $O(1)$). This is particularly relevant as in biophysics
1886: one often studies systems in stationary states while actively busy
1887: at exchanging heat with the sourroundings.
1888: %
1889: \\
1890: %
1891: \0(7) Another property, which is not as well known as it deserves, is
1892: that for hyperbolic systems, and by the Chaotic Hypothesis of
1893: Sec. 2, virtually for all chaotic evolutions, it is possible to
1894: develop a rigorous theory of coarse graining, \cite{BG97,Ga06b}.
1895: It leads to interpreting the SRB distributions as uniform
1896: distributions on the attractor; hence to a variational principle
1897: and to the existence of a Lyapunov function describing the
1898: approach to the stationary state, {\it i.e.} giving a measure of
1899: the distance from it, \cite{Ga01,Ga06}.
1900: %
1901: \\
1902: %
1903: However it also seems to lead to the conclusion that {\it entropy} of
1904: a stationary state {\it cannot be defined} if one requires that it
1905: should have properties closely analogous to the equilibrium
1906: entropy. For instance once coarse graining has been properly
1907: introduced, it is tempting to define the entropy of a stationary
1908: state as $k_B$ times the logarithm of the number of ``microcells''
1909: into which the attractor is decomposed, see Appendix A1,A2.
1910: %
1911: \\
1912: %
1913: This quantity can be used as a Lyapunov function, see \cite{Ga06}, but
1914: it depends on the size of the microcells in a nontrivial way: changing
1915: their size, the variation of the so defined entropy does not change by
1916: an additive constant depending only on the scale of the coarse
1917: graining ({\it at difference with respect to the equilibrium case}),
1918: but by a quantity that depends also on the control parameters ({\it
1919: e.g.} temperature, volume {\it etc.\ }), \cite{Ga01}.
1920: %
1921: \\
1922: %
1923: Given the interest of coarse graining, in Appendix A1 mathematical
1924: details about it are discussed in the context of the SRB distribution
1925: and CH; and a physical interpretation is presented in Appendix A2;
1926: hopefully they will also clarify the physical meaning of the two.
1927: %
1928: \\
1929: %
1930: (8) Finally it is often said that the FR should hold {\it always} or,
1931: if not, it is incorrect. In this respect it has to be stressed that
1932: the key assumption is the CH, which implies the FR {\it exactly} in
1933: time reversible situations. However it is clear that CH is an
1934: idealization and the correct attitude is to interpret deviations from
1935: FR in terms of corrections to the CH. For instance: \*
1936:
1937: \0CH implies exponential decay of time correlations. But in
1938: some cases there are physical reasons for long range time
1939: correlations.
1940: %
1941: \\
1942: %
1943: Or the CH implies that observables have values in a
1944: finite range. But there are cases in which phase space is not bounded
1945: and observables can take unbounded values (or such for practical
1946: purposes).
1947: %
1948: \\
1949: %
1950: Time reversal is necessary. But there are cases in which it
1951: is violated.
1952: %
1953: \\
1954: %
1955: The pdf of $p$ should be log-convex: but it is seldom so.
1956: \*
1957: What is interesting is that it appears that starting from CH and
1958: examining the features responsible for its violations it may be
1959: possible to compute even quantitatively the corrections to
1960: FR. Examples of such corrections already exist,
1961: \cite{CV03a,BGGZ05,Za06}. It would be interesting to have a
1962: concrete experiment, designed to test FR and try to understand the
1963: observed deviations; the BCG experiment in Sec.\ref{sec12} offers, if
1964: further developed, the possibility of simple tests making use the
1965: existing experimental apparatus and of the observations that it has
1966: proved to be accessible.
1967:
1968: \* \0{\bf Acknowledgements:} I am grateful to M. Bandi, A. Giuliani,
1969: W. Goldburg and F. Zamponi for countless comments and suggestions and
1970: to M. Bandi, W. Goldburg for providing their data, partially reported
1971: in Fig.3. Partially supported also by Institut des Hautes Etudes
1972: Scientifiques, by Institut Henri Poincar\'e and by Rutgers University. \*
1973:
1974:
1975: \section{A1: Coarse Graining, SRB and $1D$ Ising Models}
1976: \def\SEC{A1: Coarse Grain, SRB and Ising Lattice}\label{secA1}\iniz
1977:
1978: In equilibrium phase space volume is conserved and it is natural to
1979: imagine it divided into tiny ``cells'', in which all observables of
1980: interest are constant. The equilibrium distribution can be constructed
1981: simply by imagining to have divided phase space $\Si$ (``energy
1982: surface'') into cells of equal Liouville volume, small enough so that
1983: every interesting physical observable $F$ is constant in each
1984: cell. Then the dynamics is a cyclic permutation of the cells ({\it
1985: ergodic hypothesis}) so that the stationary distribution is just the
1986: volume distribution.
1987:
1988: In a way, this is an ``accident'', based on what appears to be a
1989: fundamentally incorrect premise, which leads to various difficulties
1990: as it is often considered in the context of attempts to put on firm
1991: grounds the notion of a ``coarse grained'' description of the
1992: dynamics. Confusion is also added by the simulations: the latter are
1993: sometimes interpreted as {\it de facto} coarse grained
1994: descriptions. It seems, however, essential to distinguish between
1995: coarse graining and representation of the dynamics as a permutation of
1996: small but finite cells.
1997:
1998: {\it Undoubtedly} dynamics can be represented by a permutation of small
1999: phase space volumes, as any simulation program effectively does. But it
2000: is also clear that the cells used in the simulations are {\it far too
2001: small} ({\it i.e.} of the size determined by the computer resolution,
2002: typically of double precision reals) to be identified with the coarse
2003: cells employed in phenomenological studies of statistical
2004: Mechanics.
2005:
2006: On the other hand if coarse grain cells are introduced which are not
2007: as tiny as needed in simulations the {\it dynamics will deform}
2008: them to an extent that after a short time it will no longer be
2009: possible to identify which cell has become which other cell! And this
2010: applies even to equilibrium states.
2011:
2012: In this respect it looks as an accident the fact that, nevertheless, at least
2013: in equilibrium a coarse grained representation of time evolution
2014: appears possible. And easily so, with small cells subject to the only
2015: condition of having equal volume; but the huge amount of literature on
2016: attempts at establishing a theory of coarse graining did not lead to a
2017: precise notion, nor to any agreement between different proposals.
2018:
2019: Under the CH systems are hyperbolic and a precise analysis of coarse
2020: graining seems doable, see \cite{Ga01,Ga95a} and \cite{Ga04b}. The key is
2021: that it is possible to distinguish between ``microcells'', so tiny
2022: that evolution is well approximated by a permutation on them, and
2023: ``cells'' which are still so small that the (few) interesting
2024: observables have constant value on them. The latter cells can be
2025: identified with ``coarse grain cells''; yet they are very large
2026: compared to the microcells and time evolution {\it cannot} be
2027: represented as their permutation. {\it Neither in equilibrium nor out of
2028: equilibrium.}
2029:
2030: That SRB distribution cannot be considered a permutation of naively
2031: defined coarse cells {\it seems} to be well known and to have been
2032: considered a drawback of the SRB distributions: it partly accounts for
2033: the skepticism that often, still now, accompanies them. \*
2034:
2035: {\it The point that will be made, {\rm see the review \cite{Ga04b}},
2036: is that hyperbolicity provides us with a natural definition of coarse
2037: grained cells. At the same time it tells us which is the weight to be
2038: given to each cell which, in turn, implies that each cell can be
2039: imagined containing many ``microcells'' whose evolution is a simple
2040: permutation of them {\rm(just as in numerical simulations)}.} \*
2041:
2042: In this appendix we consider for simplicity discrete time systems: in
2043: this case hyperbolic systems are described by a smooth map $S$,
2044: transitive and smoothly invertible, with the property that every phase
2045: space point $x$ is a ``saddle point''. Out of $x$ emerge the stable and the
2046: unstable manifolds $W^s(x),W^u(x)$ of complementary dimension. The
2047: expansion and contraction that take place near every point $x$ can be
2048: captured by the matrices $\partial S_u(x)$, $\partial S_s(x)$ obtained
2049: by restricting the matrix ({\it Jacobian matrix}) $\partial S(x)$, of
2050: the derivatives of $S$, to its action on the vectors tangent to the
2051: unstable and stable manifolds through $x$: the evolution $S$ maps
2052: $W^u(x)$,$W^s(x)$ to $W^u(Sx),W^s(Sx)$, respectively, and its
2053: derivative ({\it i.e.} its linearization) maps tangent vectors at $x$
2054: into tangent vectors at $Sx$.
2055:
2056: A quantitative expression of the expansion and contraction is given by
2057: the ``local expansion'' or ``local contraction'' rates defined by
2058: %
2059:
2060: \be
2061: \L^u_1(x)\defi\log|\det (\partial S)_u(x)|,\qquad
2062: \L^s_1(x)\defi-\log|\det (\partial S)_s(x)|.\label{14.1}
2063: \ee
2064: %
2065: Since time is now discrete, phase space contraction is now defined as
2066: $\s(x)=-\log|\det(\partial S)|$ and related to $\L^u_1(x),\L^s_1(x)$ by
2067:
2068: \be \s(x)=-\L_1^u(x)+\L_1^s(x)-\log\frac{\sin\d(Sx)}{\sin\d(x)},\label{14.2}\ee
2069: %
2070: where $\d(x)$ is the angle (in the metric chosen in phase space)
2071: between $W^s(x),W^u(x)$ (which is bounded away from $0$ and $\p$ by
2072: the smoothness of the hyperbolic evolution $S$).
2073:
2074: This suggests to imagine constructing a partition $\PP$ of phase space into
2075: closed regions ${\cal P}=(P_1,\ldots,P_m)$ with pairwise disjoint
2076: interiors, each of which is a ``rectangle'' defined as follows.
2077:
2078: The rectangle $P_i$, see the following Fig.5 for a visual guide,
2079: has a center $\k_i$ out of which emerge portions $C\subset
2080: W^s(\k_i),D\subset W^u(\k_i)$ of its stable and unstable manifolds,
2081: small compared to their curvature, which form the ``axes'' of $P_i$,
2082: see Fig.5. The set $P_i$, then, consists of the points $x$ obtained
2083: by taking a point $p$ in the axis $D$ and a point $q$ in the axis $C$
2084: and setting $x{\buildrel c\over =}W^s(p)\cap W^u(q)$, just as in an
2085: ordinary rectangle a point is determined by the intersection of the
2086: lines through any two points on the axes and perpendicular to them,
2087: see Fig.5. The symbol ${\buildrel c\over =}$ means that $x$ is the
2088: point closest to $p$ and to $q$ along paths in $W^s(p)$ and,
2089: respectively $W^u(q)$.\footnote{\nota This proviso is needed because
2090: often, and certainly in transitive hyperbolic maps, the full
2091: manifolds $W^s(p),W^u(q)$ are dense in phase space and intersect
2092: infinitely many times, \cite{Si72,Si77}.}
2093:
2094: Note that in a rectangle {\it anyone} of its points $\Bk$ could be the
2095: center in the above sense with a proper choice of $C,D$, so that
2096: $\k_i$ does not play a special role and essentially serves as a label
2097: identifying the rectangle. In dimension higher than $2$ the rectangles
2098: may (and will) have rather rough (non differentiable) boundaries,
2099: \cite{Bo78}.
2100:
2101: \eqfig{130}{120}{
2102: \ins{80}{85}{$C$}
2103: \ins{74}{39}{$D$}
2104: \ins{58}{69}{$\k$}
2105: \ins{120}{96}{$W^s_\g(x)$}
2106: \ins{115}{27}{$W^u_\g(x)$}
2107: \ins{22}{14}{$P$}}
2108: {fig5.eps}{}
2109:
2110: \0{\nota Fig.5: A rectangle $P$ with a pair of axes $C,D$
2111: crossing at the corresponding center $\k$.}
2112: \*
2113:
2114: It is a key property of hyperbolicity (hence of systems for which the
2115: CH can be assumed) that the partition ${\cal P}$ can be built to enjoy
2116: of a very special property.
2117:
2118: Consider the sequence, {\it history of
2119: $x$}, $\Bx(x)\defi\{\x_i\}_{i=-\infty}^\infty$ of symbols telling into
2120: which of the sets of ${\cal P}$ the point $S^ix$ is, {\it i.e.}
2121: where $x$ is found at time $i$, or $S^ix\in P_{\x_i}$. This is
2122: unambiguous aside from the zero volume set $\cal B$ of the points that
2123: in their evolution fall on the common boundary of two $P_\x$'s.
2124:
2125: Define the matrix $Q$ to be $Q_{\x,\x'}=0$, unless there is an
2126: interior point in $P_{\x}$ whose image is in the interior of
2127: $P_{\x'}$: and in the latter case set $Q_{\x,\x'}=1$. Then the history
2128: of a point $x$, which in its evolution does not visit a boundary common
2129: to two $P_\x$'s, must be a sequence $\Bx$ verifying the property,
2130: called {\it compatibility}, that, $Q_{\x_k,\x_{k+1}}=1$ for all times
2131: $k$.
2132:
2133: The matrix $Q$ tells us which sets $P_{\x'}$ can be reached from
2134: points in $P_\x$ in one time step. Then transitive hyperbolic maps
2135: admit a partition (in fact infinitely many) of phase space into
2136: rectangles ${\cal P}=(P_1,\ldots,P_m)$, so that \*
2137:
2138: \0(1) if $\Bx$ is a compatible sequence then there is a point $x$
2139: such that $S^k x\in P_{\x_k}$, see (for instance) Ch. 9 in
2140: \cite{Ga00}, (``{\it compatibility}''). The points $x$ outside
2141: the exceptional set $\cal B$ (of zero volume) determine uniquely the
2142: corresponding sequence $\Bx$.
2143:
2144: \0(2) the diameter of the set of points
2145: $E(\x_{-\frac12T},\ldots,\x_{\frac12T})$ consisting of all points
2146: which between time $-\frac12T$ and $\frac12T$ visit, in their
2147: evolution, the sets $P_{\x_i}$ is bounded above by $c\, e^{-c' T}$ for
2148: some $c,c'>0$ ({\it i.e.} the code $\V\x\to x$ determines $x$ ``{\it
2149: with exponential precision}'').
2150:
2151: \0(3) there is a power $k$ of $Q$ such that $Q^k_{\x\x'}>0$ for all
2152: $\x,\x'$ (``{\it transitivity}'').
2153: \*
2154:
2155: Hence points $x$ can be identified with sequences of symbols $\V\x$
2156: verifying the compatibility property and the sequences of symbols
2157: determine, with exponential rapidity, the point $x$ which they
2158: represent.
2159:
2160: The partitions ${\cal P}$ are called {\it Markov
2161: partitions}. Existence of ${\cal P}$ is nontrivial and rests on the
2162: chaoticity of motions: because the compatibility of all successive pairs
2163: implies that the full sequence is actually the history of a point (a
2164: clearly false statement for general partitions).\footnote{\nota The
2165: Markovian property has a geometrical meaning: imagine each $P_i$ as
2166: the ``stack'' of the connected unstable manifolds portions $\d(x)$,
2167: intersections of $P_i$ with the unstable manifolds of its points $x$,
2168: which will be called unstable ``layers'' in $P_i$. Then if
2169: $Q_{i,j}=1$, the expanding layers in each $P_i$ expand under the
2170: action of $S$ and their images {\it fully cover} the layers of $P_j$
2171: which they touch. Formally let $P_i\in\PP$ and $x\in P_i$,
2172: $\d(x){\buildrel c\over=} P_i\cap W_u(x)$: the if $Q_{i,j}=1$, {\it i.e.}
2173: if $SP_i$ visits $P_j$, it is $\d(Sx)\subset S\d(x)$.\label{footnote9}}
2174:
2175: If the map $S$ has a time reversal symmetry $I$ ({\it i.e} a smooth
2176: involution $I$, such that $IS=S^{-1}I$, see Eq.(\ref{2.1})) the
2177: partition $\PP$ can be so built that $I\PP=\PP$, hence $I
2178: P_i=P_{I(i)}$ for some $I(i)$. This is done simply by replacing $\PP$
2179: by the finer partition whose elements are $P_i\cap IP_j$, because if
2180: $\PP,\PP_1$ and $\PP_2$ are Markovian partitions also the partition
2181: $I\PP$ is such, as well as the partition $\PP_1\vee\PP_2$ formed by
2182: intersecting all pairs $P\in \PP_1$, $P'\in\PP_2$ (this is best seen
2183: from the geometric interpretation in footonote$^{\ref{footnote9}}$ and
2184: from the time reversal property that $I W_u(x)=W_s(Ix)$).
2185:
2186: A Markov partition such that $I\PP=\PP$ is called ``reversible'' and
2187: histories on it have the simple property that
2188: $(\Bx(Ix))_i=(\Bx(x))_{-I(i)}$.
2189:
2190: Markov partitions, when existing, allow us to think of the phase space
2191: points as the configurations of a ``$1$-dimensional spin system'',
2192: {\it i.e.} as sequences of finitely many symbols
2193: $\x\in\{1,2,\ldots,m\}$ subject to the ``hard core'' constraint that
2194: $Q_{\x_i,\x_{i+1}}=1$. Hence probability distributions on phase space
2195: which give $0$ probability to the boundaries of the elements of the
2196: Markov partitions (where history may be ambiguous) can be regarded as
2197: stochastic processes on the configurations of a $1$-dimensional Ising
2198: model (with finite spin $m$), and functions on phase space can be
2199: regarded as functions on the space of compatible
2200: sequences.\footnote{\nota
2201: It is worth also stressing that the ambiguity of the
2202: histories for the points which visit the boundaries of the sets of a Markovian
2203: partition is very familiar in the decimal representation of
2204: coordinates: it corresponds to the ambiguity in representing a decimal
2205: number as ending in infinitely many $0$'s or in infinitely many $9$'s.}
2206:
2207: The remarkable discovery, see reviews in \cite{Si72,Si77}, is that the
2208: SRB distribution not only can be regarded as a stochastic processes,
2209: but it {\it is a short range Gibbs distribution} if considered as a
2210: probability on the space of the compatible symbolic sequences $\Bx$ on
2211: $\PP$, and with a potential function $A(\Bx)=-\L_1^u(x( \Bx))$,
2212: see below and \cite{GBG04}.
2213:
2214: The sequences $\Bx$ are therefore much more natural, given the
2215: dynamics $S$, than the sequence of decimal digits that are normally
2216: used to identify the points $x$ via their cartesian
2217: coordinates.\footnote{\nota If the phase space points are considered
2218: as sequences $\Bx$ then the dynamics becomes a ``trivial'' left shift
2219: of histories. This happens always in symbolic dynamics, but in general
2220: it is of little interest unless compatibility can be decided by a
2221: ``hard core condition'' involving only nearest neighbors (in general
2222: compatibility is a global condition involving all symbols, {\it i.e.}
2223: as a hard core it is one with infinite range). {\it Furthermore} also
2224: the statistics of the motion becomes very well understood, because
2225: short range $1D$ Gibbs distributions are elementary and well
2226: understood.} \*
2227:
2228: \0{\bf Definition:} {\it ({\rm Coarse graining}) Given a Markovian
2229: partition $\PP$ let $\PP^T$ be the finer partition of phase space
2230: into sets of the form
2231:
2232: \be E_{\Bx}=E_{\x_{-T/2},\ldots,\x_{T/2}}\defi\bigcap_{-T/2}^{T/2} S^k
2233: P_{\x_k}.
2234: \label{14.3}\ee
2235: %
2236: The sets $E_\Bx$ will be called ``elements of a description of the
2237: microscopic states {\rm coarse grained to scale $\g$}'' if $\g$ is the
2238: largest linear dimension of the nonempty sets $E_{\Bx}$. The elements
2239: $E_{\Bx}$ of the ``coarse grained partition $\PP^T$ of phase space''
2240: are labeled by a finite string
2241: %
2242: \be \Bx\,=\, (\x_{-T/2},\ldots,\x_{T/2})\label{14.4}\ee
2243: %
2244: with $\x_i=1,\ldots,m$ and $Q_{\x_i,\x_{i+1}}=1$.}
2245: \*
2246:
2247: Define the {\it forward} and {\it backward} expansion and contraction
2248: rates as
2249:
2250: \be U^{T/2}_{u,\pm}(x)=\sum_{j=0}^{\pm T/2} \L_1^u(S^jx),\qquad
2251: U^{T/2}_{s,\pm}(x)=\sum_{j=0}^{\pm T/2} \L_1^s(S^jx)\label{14.5}\ee
2252: %
2253: and select a point $\Bk(\Bx)\in E_{\Bx}$ for each $\Bx$. Then the SRB
2254: distribution $\m_{SRB}$ and the volume distribution $\m_L$ on the
2255: phase space $\O$, which we suppose to have Liouville volume, footnote
2256: p.\pageref{2}, $V(\O)$, attribute to the {\it nonempty} sets
2257: $E_{\V\x}$ the respective probabilities $\m$ and $\m_L$
2258:
2259: \be \m(\Bx)\defi\m_{SRB}(E_{\Bx})\qquad \hbox{\rm and
2260: respectively}\qquad \m_{L}(\Bx)\defi \frac{V(E_{\Bx})}{V(\O)}\label{14.6} \ee
2261: %
2262: if $V(E)$ denotes the Liouville volume of $E$. The distributions
2263: $\m,\m_L$ are shown, \cite{GBG04,Ga00}, to be defined by
2264:
2265: \be
2266: \eqalign{\m(\Bx)\,=&\,
2267: h^T_{u,u}(\Bx)\cdot
2268: e^{\big(-U_{u,-}^{T/2}(\k(\Bx))-U^{T/2}_{u,+}(\k(\Bx))\big)}\cr
2269: \m_{L}(\Bx)\,=&\,
2270: h^T_{s,u}(\Bx)\cdot
2271: e^{\big(U_{s,-}^{T/2}(\k(\Bx))-U_{u,+}^{T/2}(\k(\Bx))\big)}\cr}\label{14.7}
2272: \ee
2273: %
2274: where $\k(\Bx)\in E_\Bx$ is the center of $P_{\x_0}$
2275: and $h^T_{u,u}(\Bx)$, $h^T_{s,u}(\Bx)$ are suitable functions of
2276: $\Bx$, {\it uniformly bounded} as $\Bx$ and $T$ vary and which are mildly
2277: dependent on $\Bx$; so that they can be regarded as constants for the
2278: purpose of the present discussion, {\it cfr.} Ch. 9 in
2279: \cite{Ga00}.
2280:
2281: If $\g$ is a scale below which all interesting observables are (for
2282: practical purposes) constant, then choosing $T=O(\log\g^{-1})$ the
2283: sets $E_\Bx$ are a coarse graining of phase space suitable for
2284: computing time averages as weighted sums over the elements of the
2285: partition.
2286:
2287: And both in equilibrium and out of equilibrium the SRB
2288: distribution {\it will not attribute equal weight} to the sets
2289: $E_\Bx$. The weight will be instead proportional to
2290: $e^{\big(-U_{u,-}^{T/2}(\k(\Bx))-U^{T/2}_{u,+}(\k(\Bx))\big)}$, {\it
2291: i.e.} to the inverse of the exponential of the expansion rate of the
2292: map $S^T$ along the unstable manifold and as a map of
2293: $S^{-\frac{T}2}\k(\Bx)$ to $S^{\frac{T}2}\k(\Bx)$. The more unstable
2294: the cells are the less weight they have. Given Eq. (\ref{14.7}) the
2295: connection with the Gibbs state with potential energy $A(\Bx)=\L_1^u(\Bx)$
2296: appears, see \cite[Sec.4.3 and Ch. 5,6]{GBG04}.
2297:
2298: The sets $E_{\Bx}$ represent macroscopic states, being just small
2299: enough so that the physically interesting observables have a constant
2300: value within them; and we would like to think that they provide us
2301: with a model for a ``{\it coarse grained}'' description of the
2302: microscopic states. The notion of coarse graining is, here,
2303: precise and, nevertheless, quite flexible because it contains a free
2304: ``resolution parameter'' $\g$. Should one decide that the resolution
2305: $\g$ is not good enough because one wants to study the system with
2306: higher accuracy then one simply chooses a smaller $\g $ (and,
2307: correspondingly, a larger $T$).
2308:
2309:
2310: \def\SEC{A2: SRB and Coarse Graining: a physicist's view}
2311: \section{A2: SRB and Coarse Graining: a physicist's view}
2312: \label{secA2}\iniz
2313:
2314: How can the analysis of Appendix A1 be reconciled with the numerical
2315: simulations, and with the naive view of motion, as a permutation of
2316: cells? The phase space volume will generally contract with time: yet
2317: we want to describe the evolution in terms of an evolution permuting
2318: microscopic states. Also because this would allow us to count the
2319: microscopic states relevant for a given stationary state of the system
2320: and possibly lead to extending to stationary nonequilibria Boltzmann's
2321: definition of entropy.
2322:
2323: Therefore we divide phase space into {\it equal} parallelepipedal {\it
2324: microcells} $\D$ of side size $\e\ll \g $ and try to discuss time
2325: evolution in terms of them: we shall call such cells ``microscopic''
2326: cells, as we do not associate them with any particular observable;
2327: they represent the highest microscopic resolution.
2328:
2329: The new microcells should be considered as realizations of objects
2330: alike to those arising in computer simulations: in simulations the
2331: cells $\D$ are the ``digitally represented'' points with coordinates
2332: given by a set of integers and the evolution $S$ is a {\it program} or
2333: {\it code} simulating the solution of equations of motion suitable for
2334: the model under study. The code operates {\it exactly} on the
2335: coordinates (the deterministic round offs, enforced by the
2336: particular computer hardware and software, should be considered part
2337: of the program).
2338:
2339: The simulation will produce (generically) a chaotic evolution ``for
2340: all practical purposes'', {\it i.e.}
2341: %
2342: \\
2343: %
2344: (1) if we only look at ``macroscopic observables'' which are constant
2345: on the coarse graining scale $\g=e^{-\frac12\lis\l T}\ell_0$ of the
2346: partition ${\cal P}^T$, where $\ell_0$ is the phase space size and
2347: $\lis\l>0$ is the least contractive line element exponent (which
2348: therefore fixes the scale of the coarse graining, by the last
2349: definition);\footnote{\nota Here it is essential that the CH holds,
2350: otherwise if the system has long time tails the analysis becomes
2351: much more incolved and so far it can be dealt, even if only
2352: qualitatively, on a case by case basis.} and
2353: %
2354: \\
2355: %
2356: (2) if we look at phenomena on time scales far shorter than the
2357: recurrence times (always finite in finite representations of motion,
2358: but of size usually so long to make the recurrence phenomenon
2359: irrelevant).\footnote{\nota To get an idea of the orders of magnitude
2360: consider a gas of $N$ particles of density $\r$ at temperature $T$:
2361: the metric on phase space will be $ds^2=\sum_i(\frac{d\V p_i^2}{k_B
2362: T}+\frac{d\V q_i^2}{\r^{-2/3}})$; hence the size of a microcell
2363: will be $\sqrt{O(N)}\,\d_0$ if $\d_0$ is the precision with which
2364: the coordinates are imagined determined (in simulations $\d_0\simeq
2365: 10^{-14}$ in double precision) as all contributions to $ds^2$ are
2366: taken of order $O(1)$. Coarse grained cells contain, in all
2367: proposals, many particles, $O(N)$, so that their size will contain a
2368: factor $\d$ rather than $\d_0$ and will be $\d/\d_0=O(N^{1/3})$
2369: larger.}
2370:
2371: The latter conclusion can be reached by realizing that \*
2372:
2373: \0(a) there has to be a small enough division into microcells that
2374: allows us to describe evolution as a map (otherwise
2375: numerical simulations would not make sense);
2376:
2377: \0(b) however the evolution map cannot be, in general, a
2378: permutation. In simulations it will happen, {\it essentially always},
2379: that it ({\it i.e.} the software program) will send two distinct
2380: microcells into the same one. It does certainly happen in
2381: nonequilibrium systems in which phase space contracts in the
2382: average;\footnote{\nota With extreme care it is sometimes, and in
2383: equilibrium, possible to represent evolution with a code which is a
2384: true permutation: the only example that I know, dealing with a
2385: physically relevant model, is in \cite{LV93}.}
2386:
2387: \0(c) even though the map will not be one-to-one, nevertheless it will
2388: be such {\it eventually}: because any map on a finite space is a
2389: permutation of the points which are recurrent. This set is the {\it
2390: attractor} of the motions, that we call $\AA$ and which will be
2391: imagined as a the collection of microcells approximating the unstable
2392: manifold and intersecting it. All such microcells will be considered
2393: taking part in the permutation: but this is not an innocent assumption
2394: and in the end is the reason why the SRB is unique, see
2395: below.
2396:
2397: \0(d) every permutation can be decomposed into cycles: each cycle will
2398: visit each coarse cell with the same frequency (unless there are more
2399: than one stationary distributions describing the asymptotics of a set
2400: of microcells initially distributed uniformly, a case that we exclude
2401: because of the transitivity assumption). Hence it is not restrictive to
2402: suppose that there is only one cycle (``ergodicity'' on the
2403: attractor). \*
2404:
2405: {\it Then} consistency between the expansion of the unstable
2406: directions and the existence of a cyclic permutation of the microcells
2407: in the attractor $\AA$ {\it demands that the number of microcells in
2408: each coarse grained cell $E_\Bx$, Eq.(\ref{14.3}), must be inversely
2409: proportional to the expansion rate}, {\it i.e.} it has to be given by
2410: the first of Eq. (\ref{14.7}).
2411:
2412:
2413: \eqfig{200}{60}{
2414: \ins{-20}{40}{$\EE(\Bx)$}}{cell.eps}{}
2415:
2416: \0{\nota Fig.4: A very schematic and idealized drawing of the
2417: attractor layers $\st\D(\Bx)$, remaining after a transient time,
2418: inside a coarse cell $\st\EE(\Bx)$; the second drawing (indicated by
2419: the arrow) represents schematically what the layers really are, if
2420: looked closely: namely collections of microcells laying uniformly on
2421: the attractor layers, {\it i.e.} the discretized attractor
2422: intersected with the coarse cell.} \*
2423: %
2424: More precisely we imagine, developing a heuristic argument, that
2425: the attractor in each coarse cell $\EE(\Bx)$ will appear as a stack of
2426: a few portions of unstable manifolds, the ``layers'' of
2427: footnote$^{\ref{footnote9}}$, whose union form the (disconnected)
2428: surface $\D(\Bx)$ intersection between $\EE(\Bx)$ and the
2429: attractor. Below $\D(\Bx)$ will be used to denote both the set and its
2430: surface, as the context demands. The stack of connected surfaces
2431: $\D(\Bx)$ is imagined covered uniformly by $N(\Bx)$ microcells, see Fig.4.
2432:
2433: Let $t\defi T+1$. Transitivity implies that there is a smallest
2434: integer $m\ge0$ such that $S^{t+m}\EE(\Bx)$ intersects all other
2435: $\EE(\Bx')$: the integer $m$ is $t$-independent (and equal to the
2436: minimum $m$ such that $Q^m_{\s,\s'}>0$). In $t+m$ time steps each
2437: coarse cell will have visitied all the others and the layers
2438: describing the approximate attractor in a single coarse cell will have
2439: been expanded {\it to cover the entire attractor} for the map
2440: $S^{t+m}$.%
2441: \footnote{\nota To see this it is convenient to remark that the
2442: $S^{t+m}$-image of a layer $\d(x)\subset \D(\Bx)$ of the attractor
2443: will cover some of the layers of $\D(\Bx)$, because $S^t\EE(\Bx)$
2444: visits and fully covers all coarse cells $\EE(\Bx')$, see
2445: footnote$^{\ref{footnote9}}$. Hence $S^{t+m}\D(\Bx)$ will fully cover
2446: at least part of the layers of the attractor in $\EE(\Bx)$. Actually
2447: {\it it will cover the whole of} $\D(\Bx)$, because if a layer of
2448: $\D(\Bx)$ was left out then it will be left out of all the iterates of
2449: $S^{t+m}$ and a nontrivial invariant subset of the attractor for $S^t$
2450: would exist.} The latter coincides with the attractor for $S$ because
2451: $S^j$ is transitive for all $j$ if it is such for $j=1$ and this
2452: property has to be reflected by the discretized dynamics at least if
2453: $j$ is very small compared to the (enormous) recurrence time on the
2454: discrete attractor as $t$ is, being a time on the coarse grain scale.
2455: %and coincides with it if
2456: %$t+m$ is prime with respect to the (enormous) recurrence times, which
2457: %will be supposed for simplicity and without further discussing a
2458: %condition that can be eliminated.
2459:
2460: Suppose first that $m=0$, hence $S^t\D(\Bx)$ is the entire attractor
2461: for all $\Bx$. This is an assumption useful to exhibit the idea but
2462: unrealistic for invertible maps: basically this is realized in the
2463: closely related SRB theory for a class of non invertible expansive
2464: maps of the unit interval).
2465:
2466: So the density of microcells will be $\r(\Bx)=\frac{N(\Bx)}{\D(\Bx)}$
2467: and under time evolution $S^t$ the unstable layers $\D(\Bx')$ in
2468: $\EE(\Bx)$ expand and cover all the layers in the cells $\EE(\Bx')$.
2469: %
2470: If the coarse cell $\EE(\Bx)$ is visited, in $t=T+1$ time steps, by
2471: points in the coarse cells $\Bx'$, a property that will be
2472: symbolically denoted $\EE_{\Bx'}\in S^{-t}\EE(\Bx)$, a fraction
2473: $\n_{\Bx,\Bx'}$ of the $N(\Bx')$ microcells will end in the coarse
2474: cell $\EE(\Bx)$, and $\sum_{\Bx}\n_{\Bx,\Bx'}=1$. Then consistency
2475: with evolution as a cyclic permutation demands
2476:
2477: \be N(\Bx)=\sum_{\Bx'}
2478: \frac{N(\Bx')}{\D(\Bx')}\frac1{e^{\L_{u,T}(\Bx')}}\D(\Bx)
2479: \defi\LL(N)(\Bx),\qquad\hbox{\it i.e.}
2480: \label{15.1}\ee
2481: %
2482: because the density of the microcells on the images of $\D(\Bx')$
2483: decreases by the expansion factor $e^{\L_{u,T}(\Bx')}$, so that
2484: $\n_{\Bx,\Bx'}= \frac{\D(\Bx)}{\D(\Bx')}\frac1{e^{\L_{u,T}(\Bx')}}$.
2485:
2486: As a side remark it is interesting to point out that for the density
2487: $\r(\Bx)$ Eq.(\ref{15.1}) becomes simply $\r(\Bx)=\sum_{\Bx'}
2488: e^{-\L_{u,T}(\Bx')}\r(\Bx')$, closely related to the similar equation
2489: for invariant densities of Markovian surjectiive maps of the unit
2490: interval, \cite{GBG04}.
2491:
2492: The matrix $\LL$ has all elements $>0$ (because $m=0$) and therefore
2493: has a simple eigenvector $v$ with positive components to which
2494: corresponds the eigenvalue $\l$ with maximum modulus: $v=\l\,\LL(v)$
2495: (the ``Perron-Frobenius theorem'') with $\l=1$ (because
2496: $\sum_{\Bx}\n_{\Bx,\Bx'}=1$). It follows that the consistency
2497: requirement uniquely determines $N(\Bx)$ as proportional to $v_\Bx$.
2498: Furthermore $S^{t}\D(\Bx)$ is the entire attractor; then its surface
2499: is $\Bx$ independent and equal to $e^{\L_{u,T}(\Bx)}\D(\Bx)$:
2500: therefore $N(\Bx)=\,const\, e^{-\L_{u,T}(\Bx)}$.
2501:
2502: The general case is discussed by considering $S^{t+m}$ instead of
2503: $S^{t}$: this requires taking advantage of the properties of the
2504: ratios $e^{\L_{u,T}(\Bx)}/e^{\L_{u,T+m}(\Bx)}$. Which are not only
2505: uniformly bounded in $T$ but also only dependent on the sequence
2506: $\Bx=(\x_{-\frac12T},\ldots,\x_{\frac12T})$ through a few symbols with labels
2507: near $-\frac12T$ and $\frac12T$: this correction can be considered part of the
2508: factors $h^T_{u,u}$ in the rigorous formula Eq.(\ref{14.7}).
2509:
2510:
2511: Note that $e^{\L_{u,T}(\Bx)}\D(\Bx)=$ constant reflects Pesin's
2512: formula, \cite{GBG04}, for the approximate dynamics considered here.
2513: \*
2514:
2515: So the SRB distribution arises naturally from assuming that dynamics
2516: can be discretized on a regular array of point (``microcells'') and
2517: become a one cycle permutation of the microcells on the
2518: attractor. This is so under the CH and holds whether the dynamics is
2519: conservative (Hamiltonian) or dissipative.
2520:
2521: \*
2522:
2523: \0{\it Remark:} It is well known that hyperbolic systems admit
2524: (uncountably) many invariant probability distributions, besides the
2525: SRB. This can be seen by noting that the space of the configurations
2526: is identified with a space of compatible sequences. On such a space
2527: one can define uncountably many stochastic processes, for instance by
2528: assigning an arbitrary short range translation invariant potential,
2529: and regarding the corresponding Gibbs state as a probability
2530: distribution on phase space. However the analysis just presented
2531: apparently singles out SRB as the unique invariant distribution. This
2532: is due to our assumption that, in the discretization, microcells are
2533: regularly spaced and centered on a regular discrete lattice and
2534: evolution eventually permutes them in a (single, by transitivity)
2535: cycle consisting of the microcells located on the attractor (and
2536: therefore locally evenly spaced, as inherited from the regularity of
2537: the phase space discretization).
2538:
2539: \0Other invariant distributions can be obtained by custom made
2540: discretizations of phase space which will not cover the attractor in a
2541: regular way. This is what is done when other distributions, ``not
2542: absolutely continuous with respect to the phase space volume'', are to
2543: be studied in simulations. A paradigmatic example is given by the map
2544: $x\to 3x \, {\rm mod}\, 1$: it has an invariant distribution
2545: attributing zero probability to the points $x$ that, in base $3$, lack
2546: the digit $2$: to find it one has to write a program in which data
2547: have this property and make sure that the round off errors will not
2548: destroy it. Almost any ``naive'' code that simulates this dynamics using double
2549: precision reals represented in base $2$ will generate, instead, the
2550: corresponding SRB distribution which is simply the Lebesgue measure on
2551: the unit interval (which is the Bernoulli process on the symbolic
2552: dynamics giving equal probability $\frac13$ to each digit).
2553:
2554: \*
2555:
2556: The physical representation of the SRB distribution just obbtained, see
2557: \cite{Ga95a,Ga00}, shows that there is no conceptual difference
2558: between stationary states in equilibrium and out of equilibrium. In
2559: both cases, if motions are chaotic they are permutations of
2560: microcells and the {\it SRB distribution is simply equidistribution
2561: over the recurrent microcells}. In equilibrium this gives the Gibbs
2562: microcanonical distribution and out of equilibrium it gives the SRB
2563: distribution (of which the Gibbs one is a very special case).
2564:
2565: The above heuristic argument is an interpretation of the mathematical
2566: proofs behind the SRB distribution which can be found in
2567: \cite{Bo75,GBG04}, (and heuristically is a proof in itself). Once
2568: Eq. (\ref{14.7}) is given, the expectation values of the observables
2569: in the SRB distributions can be formally written as sums over suitably
2570: small coarse cells and symmetry properties inherited from symmetries
2571: of the dynamic become transparent. The Fluctuation Theorem is a simple
2572: consequence of Eq. (\ref{14.7}), see Appendix A3: however it is
2573: conceptually interesting because of the surprising unification of
2574: equilibrium and nonequilibrium behind it.
2575:
2576: %\vfill\eject
2577: The discrete repesentation, in terms of coarse grain cells and
2578: microcells leads to the possibility of counting the number $\NN$ of
2579: the microcells and therefore to define a kind of entropy function: see
2580: \cite{Ga01} where the detailed analysis of the counting is performed
2581: and the difficulties arising in defining an entropy function as
2582: $k_B\log \NN$ are critically examined.
2583:
2584: \section{A3: Why does FT hold?}
2585: \def\SEC{A3: Why FT?}\label{secA3}\iniz
2586:
2587: As mentioned the proof of FT in quite simple, \cite{GC95}. By the
2588: first of Eq. (\ref{14.5}), (\ref{14.7}) and by the theory of
2589: $1D$-short range Ising models, see \cite{Ga95b} for details, the
2590: probability that $p$ is in a small interval centered at $p$ compared
2591: to the probability that it is in the opposite interval is
2592:
2593: \be
2594: \frac{P_\t(p)}{P_\t(-p)}=\frac{\sum_{i \to
2595: p\s_+\t}e^{-\sum_{-\t/2}^{\t/2} \L^u_{1}(S^k \k_i)+B(i,\t)}}
2596: {\sum_{i \to
2597: -p\s_+\t}e^{-\sum_{-\t/2}^{\t/2} \L^u_{1}(S^k \k_i) +B(i,\t)}}
2598: \label{16.1}\ee
2599: %
2600: where $\sum_{i \to p\s_+\t}$ is sum over the centers $\k_i$ of the
2601: rectangles $E_i$ labeled by $i\defi(\x_{-\t/2},\ldots,\x_{\t/2})$ with
2602: the property
2603:
2604: \be \sum_{k=-\t/2}^{\t/2}
2605: \s(S^k\k_i)+B(i,\t)\simeq p\s_+\t\,\label{16.2}\ee
2606: %
2607: where $\simeq$ means that the left hand side is contained in a very
2608: small interval (of size of order $O(1)$, \cite{Ga95b}, call it $b$)
2609: centered at $p\s_+\t$; the $B(i,\t)$ is a term of order $1$ (a
2610: boundary term in the language of the Ising model interpretation of the
2611: SRB distribution): $|B(i,\t)|\le b<+\infty$: and it takes also into
2612: account the adjustments to be made because of the
2613: arbitrariness of the choice of $\k_i$.%
2614: \footnote{Which is taken here $\Bk_i=\,$the center of $P_{\x_0}$, but
2615: which could equivalently made by choosing other points in $E_\Bx$,
2616: for instance by continuing the string
2617: $i=(\x_{-\t/2},\ldots,\x_{\t/2})$ to the right and to the left,
2618: according to an a priori fixed rule depending only on $\x_{\t/2}$
2619: and $\x_{-\t/2}$ respectively. Thus turning it to a biinfinite
2620: compatible string $\Bx_i$ which therefore fixes a new point
2621: $\Bk'_i$.} Independence on $i,\t$ of the bound on $B(i,\t)$
2622: reflects smoothness of $S$ and elementary properties of
2623: short range $1D$ Ising chains, \cite{Ga95b}.
2624:
2625: Suppose that the symbolic dynamics has been chosen time reversible,
2626: {\it i.e.} the time reversal map $I$ maps $P_i$ into $IP_i=P_{I(i)}$
2627: for some $I(i)$: this is not a restriction as discussed in Appendix A1.
2628: Then the above ratio of sums can be rewritten as a ratio of sums over
2629: the same set of labels,
2630:
2631: \be
2632: \frac{P_\t(p)}{P_\t(-p)}=\frac{\sum_{i \to
2633: p\s_+\t}e^{-\sum_{-\t/2}^{\t/2} \L^u_{1}(S^k \k_i)+B(i,\t)}}
2634: {\sum_{i \to
2635: p\s_+\t}e^{-\sum_{-\t/2}^{\t/2} \L^u_{1}(S^k I(\k_i)) +B(I(i),\t)}}.
2636: \label{16.3}\ee
2637: %
2638: Remark that $\L^u_1(Ix)=-\L^s_1(x)$ (by time reversal symmetry) and
2639: that (by Eq. (\ref{14.3})) $ \sum_{k=-\t/2}^{\t/2}
2640: (\L^u_1(S^k(x))+\L^s_1(S^{-k}(x)))$ can be written as
2641:
2642: \be
2643: \sum_{k=-\t/2}^{\t/2}
2644: (\L^u_1(S^k(x))+\L^s_1(S^{k}(x)))=\sum_{k=-\t/2}^{\t/2}
2645: \s(S^kx)+B(x,\t)\label{16.4}\ee
2646: %
2647: with $B(x,\t)\le b$ (again by the smoothness of $S$), possibly
2648: redefining $b$.
2649:
2650: Therefore the ratio of corresponding terms in the numerator and
2651: denominator ({\it i.e.} terms bearing the same summation label $i$) is
2652: precisely $p\s_+\t$ up to $\pm3b$. Hence
2653:
2654: \be e^{\t\s_+\,p\,-3b}\le \frac{P_\t(p)}{P_\t(-p)}< e^{\t\s_+\,p\,+3b}
2655: \label{16.5}\ee
2656: %
2657: so that FT holds for finite $\t$ with an error $\pm\frac{3b}\t$,
2658: infinitesimal as $\t\to+\infty$. For a detailed discussion of the
2659: error bounds see \cite{Ga95b}.
2660:
2661: {\it Of course} for all this to make sense the value of $p$ must be
2662: among those which not only are possible but also such that the values
2663: close enough to possible values are possible. This means that $p$ has
2664: to be an internal point to an interval of values that contains limit
2665: points of $\lim_{\t\to+\infty} \frac1\t\sum_{k=0}^{\t} \frac{\s(S^k
2666: x)}{\s_+}$ for a set of $x$'s with positive SRB probability: the value
2667: $p^*$ in FT is the supremum among the value of $p$ with this property,
2668: \cite{Ga95b} (contrary to statements in the literature this physically
2669: obvious remark is explicitly present in the original papers: and one
2670: should not consider the three contemporary references,
2671: \cite{GC95,Ga95b,GC95b}, has having been influenced by the doubts on
2672: this point raised much later.)
2673:
2674: The assumptions have been: (a) existence of a Markovian partition,
2675: {\it i.e.} the possibility of a well controlled symbolic dynamics
2676: representation of the motion; (b) smooth evolution $S$ and (c) smooth
2677: time reversal symmetry: the properties (a),(b) are equivalent to the
2678: CH. Of course positivity of $\s_+$ is {\it essential}, in spite of
2679: contrary statements; if $\s_+=0$ the leading terms would come
2680: from what has been bounded in the remainder terms and, in any event
2681: the analysis world be trivial, with or without chaoticity assumptions,
2682: \cite{BGGZ05}.
2683:
2684: Since Lorenz, \cite{Lo63}, symbolic dynamics is employed to represent
2685: chaos and many simulations make currently use of it; smoothness has
2686: always been supposed in studying natural phenomena (lack of it being
2687: interpreted as a sign of breakdown of the theory and of necessity of a
2688: more accurate one); time reversal is a fundamental symmetry of nature
2689: (realized as $T$ or $TCP$ in the Physics notations). Hence in spite of
2690: the ease in exhibiting examples of systems which are not smooth, not
2691: hyperbolic, not time reversal symmetric (or any subset thereof) the CH
2692: still seems a good guide to understand chaos.
2693:
2694: \section{A4: Harmonic Thermostats}
2695: \def\SEC{A4: Harmonic Thermostats}\label{secA4}\iniz
2696:
2697:
2698: Here the ``efficiency'' of a harmonic thermostat is discussed. It
2699: turns out that in general a thermostat consisting of infinite free
2700: systems is a very simple kind of Hamiltonian thermostat, but it has to
2701: be considered with caution as it can be inefficient in the sense that
2702: it might not drive a system towards equilibrium (i.e. towards a Gibbs
2703: distribution). In the example given below a system in interaction with
2704: an infinite harmonic reservoir at inverse temperature $\b$ is
2705: considered.
2706: %
2707: It is shown that the interaction can lead to a stationary state, of
2708: the system plus reservoir, which is not the Gibbs state at temperature
2709: $\b^{-1}$. The following is a repetition of the analysis in
2710: \cite{ABGM72}, adapting it to the situation considered here.
2711:
2712: A simple model is a $1$-dimensional harmonic oscillators chain, of
2713: bosons or fermions, initially in a Gibbs state at temperature
2714: $\b^{-1}$. The Hamiltonian for the equilibrium {\it initial} state
2715: will be
2716:
2717: \be H_0=\sum_{x=1}^{N-1}
2718: -\frac{\hbar^2}{2m}\D_{q_x}+\sum_{x=1}^{N-1}\frac{m\o^2}2
2719: q_x^2+\sum_{x=1}^N\frac{m\m^2}2 (q_x-q_{x-1})^2\label{17.1}\ee
2720: %
2721: with boundary conditions $q_0=q_N=0$ and $\hbar,m,\o^2,\m^2>0$. The
2722: initial state will be supposed to have a density matrix
2723: $\r_0=\frac{e^{-\b H_0}}{{\rm Tr}\, e^{-\b H_0}}$. Time evolution will
2724: be governed by a {\it different} Hamiltonian
2725:
2726: \be H_\l=H_0+\frac{m\l}2 q_1^2,\qquad \l+\o^2>0\label{17.2}\ee
2727: %
2728: The question of ``thermostat efficiency'' is: does $\r_t\defi
2729: e^{\frac{i}{\hbar} t H_\l}\r_0 e^{-\frac{i}{\hbar} t H_\l}$ converge
2730: as $t\to+\infty$ to $\r_\infty=\frac{e^{-\b H_\l}}{{\rm Tr}\, e^{-\b
2731: H_\l}}$. Or: does the system consisting in the oscillators labeled
2732: $2,3,\ldots$ succeed in bringing up to the new equilibrium state
2733: the oscillator labeled $1$?
2734: %
2735: Convergence means that the limit $\media{A}_{\r_t}\tende{t\to+\infty}$
2736: $\media{A}_{\r_\infty}$ exists, at least for the observables $A$
2737: essentially localized in a finite region.
2738:
2739: The Hamiltonian in Eq.(\ref{17.2}) can be diagonalized by studying the
2740: matrix
2741:
2742: \be
2743: V_\l=m\,\pmatrix{\o^2+2 \m^2+\l& -\m^2&0 &\ldots\cr
2744: -\m^2 &\o^2+ 2\m^2 &-\m^2 &\ldots\cr
2745: 0 &-\m^2 &\o^2+2 \m^2&\ldots\cr
2746: \ldots&\ldots&\ldots&\ldots}\defi V_0+\l m P_1\label{17.3}\ee
2747: %
2748: The normalized eigenstates and respective eigenvalues of $V_0$ are
2749:
2750: \be
2751: \Ps^0_k(x)\defi\sqrt{\frac{2}{N}}\,\sin\frac{\p k}Nx,\qquad
2752: \L^0_{k}=m\,\Big(\o^2+{2\m^2(1-\cos\frac{\p k}N)}\Big)\label{17.4}\ee
2753: %
2754: and the vectors $\Ps^0_k$ will be also denoted $\ket{k}$ or $\ket{\Ps^0_k}$.
2755:
2756: To solve the characteristic equation for $V_\l$, call $\Ps$ a generic
2757: normalized eigenvector with eigenvalue $\L$; the eigenvalue equation is
2758:
2759: \be \bra{k}\ket\Ps (\L^0_{k}-\L)+\l\,m\,\bra k
2760: \ket\BO\,\bra\BO\ket\Ps=0\label{17.5}\ee
2761: %
2762: where $\BO$ is the vector $\BO=(1,0,\ldots,0)\in\CC^{N-1}$, so that
2763: $P_1=\ket\BO \brav\BO$. Hence, noting that $\bra{\BO}\ket{\Ps}$ cannot
2764: be $0$ because this would imply that $\L=\L^0_{k}$ for some $k$ and
2765: therefore $\ket{\Psi}=\ket{k}$ which contradicts $\bra\BO\ket\Ps=0$,
2766: it is
2767:
2768: \be \bra{k}\ket\Ps=-\l\,m\,
2769: \frac{\bra{k}\ket\BO\cdot\bra\BO\ket\Ps}{\L^0_{k}-\L}\label{17.6}\ee
2770: %
2771: and the compatibitity condition that has to be satisfied is
2772:
2773: \be
2774: \frac{\bra\BO\ket\Ps}{\l\,m}\,=
2775: \,\sum_{k=1}^{N-1}\frac{|\bra\BO\ket{k}|^2}{\L-\L^0_{k}}
2776: \bra\BO\ket\Ps=
2777: \,\sum_{k=1}^{N-1}\frac{2\sin^2\frac{\p
2778: k}{N}}{N}\frac{\bra\BO\ket\Ps}{\L-\L^0_{k}}.\label{17.7}\ee
2779: %
2780: Once Eq.(\ref{17.7}) is satisfied, Eq.(\ref{17.6}) imply that
2781: the eigenvalue equation, Eq.(\ref{17.5}), is satisfied, and by a
2782: $\ket\Psi\ne0$ (determined up to a factor).
2783:
2784: The Eq.(\ref{17.7}) has $N-1$ solutions, corresponding to the $N-1$
2785: eigenvalues of $V_\l$. This follows by comparing the graph of
2786: $y(\L)\equiv\frac1{\l m}$ with the graph of the function of $\L$ in the
2787: intermediate term of Eq.(\ref{17.7}). One of the solutions remains
2788: isolated in the limit $N\to\infty$, because the equation
2789:
2790: \be 1=\frac{2\,\l\,m}{\p}\int_0^\p\frac{\sin^2\k}{\L-\L^0(\k)}
2791: d\k,\qquad \L^0(\k)\defi
2792: m\,\Big(\o^2+4\m^2\sin^2\frac{\k}2\Big)\label{17.8}\ee
2793: %
2794: has, uniformly in $N$, only one isolated solution for $\L<\inf \L^0(\k)=m
2795: \o^2$ if $\l<0$, or for $\L>\sup \L^0(\k)$ if $\l<0$. Suppose for
2796: definiteness that $\l<0$.
2797:
2798: Let $\Ps^\l_k(x),\,k=1,\ldots,N-1$, be the corresponding
2799: eigenfunctions. The matrices $ U_{\l;k,x}=\Ps^\l_k(x)$ are unitary and
2800: $(U_\l)_{\l=0}\equiv U_0$. It is $U_{0;k,x}=\sqrt{\frac2{N}}\sin
2801: \frac{\p k}N x$ and
2802: $\bra{\Ps^0_k}\ket{\Ps^\l_{k'}}=\frac{\bra{k}\ket\BO}{Z_N(k')(\L^\l_{k'}-\L^0_k)}$
2803: with $Z_N(k')^2=\sum_{k}
2804: \frac{|\bra{k}\ket\BO|^2}{(\L^\l_{k'}-\L^0_k)^2}$ by Eq.(\ref{17.6}).
2805: %
2806: Then setting $\a^\pm_x= \frac{p_x\pm
2807: iq_x}{\sqrt2}$ let
2808:
2809: \be
2810: a^+_{\l;k} \defi (U_\l \Ba^+)_k, \qquad a^-_{\l;k}=\defi
2811: (\Ba^-U_\l^*)_k
2812: \label{17.9}\ee
2813: %
2814: where $U^*$ is the adjoint of $U$ (so that $UU^*=1$ if $U$ is unitary).
2815: It is
2816:
2817: \be \a^+_{x}=\sum_k \lis U_{\l;k,x} a^+_{\l;k},\qquad a^+_{\l;k }=
2818: \sum_{h,y} U_{\l;k,y}\lis U_{0;h,y} a^+_{0;h}
2819: \label{17.10}\ee
2820: %
2821: if the overbars denote complex conjugation.
2822:
2823: The operators $a^\pm_{\l,k}$ will be creation and
2824: annihilation operators for quanta with energy $\hbar
2825: \sqrt\frac{\L^\l_k}{m}\defi E_\l(k)$. So a state with $n_k=0,1,\ldots$
2826: quanta in state $k$ will have energy $\sum_k E_\l(k)(n_k+\frac12)$.
2827:
2828: Consider the observable $a^+_{\l,1}a^-_{\l,1}=A$. Its average is
2829: {\it time independent}, in the evolution generated by $H_\l$, and if
2830: $W\defi U_\l U^*_0$ it is equal to
2831:
2832: \be
2833: \eqalign{&\media{A}_{\r_t}\equiv \media{A}_{\r_0}\equiv {\,\rm Tr\,}
2834: \r_0 (W \V a^+_0)_1 (W \V a^-_0)_1\cr
2835: &=\sum_k {\rm Tr\,}\r_0W_{1,k} W_{1,k'} a^+_{0,k}a^-_{0,k'}
2836: =\sum_{k=1}^{N-1}|W_{1,k}|^2 \frac{\sum_{n=0}^{n_f} e^{-\b E_0(k)n}\,n}
2837: {\sum_{n=0}^{n_f} e^{-\b E_0(k)n}}
2838: \cr}\label{17.11}\ee
2839: %
2840: where $n_f=1$ if the statistics of the quanta is fermionic (this was
2841: the case in \cite{ABGM72}) or $n_f=+\infty$ if it is bosonic. In the
2842: two cases the result is
2843:
2844: \be \sum_k|W_{1,k}|^2\frac1{e^{\b E_0(k)}\pm1}\label{17.12}\ee
2845: %
2846: If the system reached thermal equilibrium, setting
2847: $\r_\l(k)\defi\frac1{e^{\b E_\l(k)}\pm1}$, this should be $\r_\l(1)$,
2848: which is impossible, as it can be checked by letting $\b\to+\infty$
2849: and remarking that it is $E_\l(1)<E_0(1)$ with a difference positive
2850: uniformly in $N$.
2851: %
2852: Furthermore the observable $A$ is localized near the site $x=1$:
2853: because the wave function of the lowest eigenvalue is
2854: $\frac1{Z_N(1)}\sum_h
2855: \frac{\bra{h}\ket\BO}{\L^0_h-\L^\l_k}\ket{\Ps^0_h}$ so that
2856:
2857: \be\Ps^\l_1(x)=\frac1{Z_N(1)}\sum_h\frac{\Ps^0_h(1)\Ps^0_h(x)}{\L^\l_1-\L^0_h}
2858: \tende{N\to\infty}\frac 1{Z_\infty}
2859: \frac2\p\int_0^\p\frac{\sin\k\,\sin\k
2860: x}{\L^\l_1-\L^0(\k)}\,d\k\label{17.13}
2861: \ee
2862: %
2863: and the integral tends to $0$ as $x\to\infty$ faster than any
2864: power, so that $0<Z_\infty<\infty$
2865: and $\Ps^\l_1$ is normalizable.
2866:
2867: Therefore the thermostatic action of the system in
2868: the sites $2,3,\ldots$ on the site $1$ is {\it not efficient} and the
2869: state does not evolve towards the Gibbs state at temperature
2870: $\b^{-1}$, not even in the limit $N\to+\infty$.
2871:
2872: This result should be contrasted with the closely related case in
2873: which the system oscillator at $1$ plus the others is started in a
2874: equilibrium state for $H_\l$ and at time $0$ is evolved with
2875: Hamiltonian $H_0$. In this case the system thermalizes properly, see
2876: the analogous analysis in \cite{ABGM72}, see also \cite{HL73b} for a
2877: large class of related examples.
2878:
2879: Of course the question of effectiveness of a thermostat could be
2880: discussed also for non linear theormostats, finite or infinite. It
2881: seems that, under mild assumptions, non linear thermostat models should
2882: be efficient, {\it i.e.} generate proper heat exchanges even when
2883: acting only at the boundary as in the case of the thermostats
2884: considered in Sec.\ref{sec9}. The analysis in \cite{GG07} gives some
2885: preliminary evidence in this direction.
2886:
2887: Harmonic thermostats are nevertheless very interesting, provided the
2888: above pathologies are excluded by a careful formulation of the models:
2889: see for instance \cite{HL73b}, see also \cite{HI05}. It is also clear
2890: that the pathologies seem to be related to the fact that the
2891: thermostats constituents are ``not interacting'' or ``linearly
2892: interacting'': their origin in the above analysis is shown to be
2893: related to the existence of isolated eigenvalues of the Hamiltonian at
2894: the bottom of the spectrum and this is the property that should be
2895: excluded. The pathologies are likely to be absent in models in which
2896: there is nonlinear interaction within the thermostats constituents so
2897: that such models should be perfectly well behavng ({\it i.e.}
2898: efficient in the sense of this paper). However the latter models are
2899: also highly nontrivial even at a purely mathematical level.
2900:
2901:
2902: \def\SEC{A5: Bohmian Quantum Systems}
2903: \section{A5: Bohmian Quantum Systems}
2904: \label{secA5}\iniz
2905:
2906: Consider the system in Fig.1 and suppose, as in Sec.\ref{sec10},
2907: that the nonconservative force $\V E(\V X_0)$ acting on the system
2908: vanishes, {\it i.e.} consider the problem of heat flow through
2909: $\CC_0$. Let $H$ be the operator on $L_2(\CC_0^{3N_0})$, space of
2910: symmetric or antisymmetric wave functions $\Ps$,
2911:
2912: \be H_{\V X}=
2913: -\frac{\hbar^2}{2m}\D_{\V X_0}+ U_0(\V X_0)+\sum_{j>0}\big(U_{0j}(\V X_0,\V
2914: X_j)+U_j(\V X_j)+K_j\big)\label{18.1}\ee
2915: %
2916: where $\D_{\V X_0}$ is the Laplacian, and note that its spectrum
2917: consists of eigenvalues $E_n=E_n(\{\V X_j\}_{j>0})$, depending on the
2918: configuration $\V X\defi\{\V X_j\}_{j>0}$,
2919:
2920: Thermostats will be modeled as assemblies of classical particles as
2921: in Sec.\ref{sec9}: thus their temperature can be defined as the
2922: average kinetic energy of their particles and the question
2923: of how to define it does not arise.
2924:
2925: The viewpoint of Bohm on quantum theory seems quite well adapted to
2926: the kind of systems considered here. A system--reservoirs model
2927: can be the {\it dynamical system} on the variables
2928: $\big(\Ps,\V X_0,(\{\V X_j\},$ $\{\V{{\dot X}}_j\})_{j>0}\big)$
2929: defined by
2930:
2931: \be \eqalign{
2932: -i\hbar {\dot\Ps(\V X_0)}=& \,( H_{\V X}\Ps)(\V X_0),\cr
2933: \V{{\dot X}}_0=&\,\hbar\,%
2934: \Im\,\,\frac{\BDpr_{\V X_0} \Ps(\V X_0)}{\Ps(\V X_0)},
2935: \kern4mm{\rm and\ for}\
2936: j>0\cr
2937: \V{{\ddot X}}_j=&-\Big(\partial_j U_j(\V X_j)+
2938: \partial_j U_j(\V X_0,\V X_j)\Big)-\a_j \V{{\dot X}}_j\cr
2939: \a_j\defi&\frac{W_j-\dot U_j}{2 K_j}, \qquad
2940: W_j\defi -\V{{\dot X}}_j\cdot \V\partial_j U_{0j}(\V X_0,\V
2941: X_j)\cr
2942: }\label{18.2} \ee
2943: %
2944: here the first equation is Schr\"odinger's equation, the second is the
2945: vlocity of the Bohmian particles carried by the wave $\Ps$, the others are
2946: equations of motion for the thermostats particles analogous to the one
2947: in Eq.(\ref{9.1}), (whose notation for the particles labels is
2948: adopted here too). Evolution maintains the thermostats kinetic
2949: energies $K_j\equiv \frac12\V{{\dot X}}_j^2$ exactly constant so that
2950: they will be used to define the thermostats temperatures $T_j$ via
2951: $K_j=\frac32 k_B T_j N_j$, as in the classical case.
2952:
2953: Note that if there is no coupling between system and thermostats, {\it
2954: i.e.} the system is ``isolated'', then there are many invariant
2955: distributions: {\it e.g.} the probability distributions $\m$
2956: proportional to
2957:
2958: \be
2959: \sum_{n=1}^\infty e^{-\b_0
2960: E_n}\d(\Ps-\Ps_n\,e^{i\f_n})\,\big|\Ps(\V X_0)|^2
2961: {d\f_n}d\V X_0\,\prod_j\d(\dot{\V X}_j^2-2K_j) d\dot{{\V
2962: X_j}}d\V X_j
2963: \label{18.3}\ee
2964: %
2965: where $E_n$ and $\Ps_n$ are time independent, under the assumed
2966: absence of interaction between system and thermostats, and are the
2967: eigenvalues and the correspoding eigenvectors of $H$. Then the
2968: distributions $\m$ are invariant under the time evolution.
2969:
2970: Time invariance of this kind of distributions is discussed in
2971: \cite[Sec.4]{DGZ92}, where it appears as an instance of what is called
2972: there a ``quantum equilibrium''. The average value of an observable
2973: $O(\V X_0)$, which depends only on position $\V X_0$, will be the
2974: ``usual'' Gibbs average
2975:
2976: \be \media{O}_\m=Z^{-1}\int {\rm Tr}\,(e^{-\b_0 H}O))\label{18.4}\ee
2977: %
2978:
2979: For studying nonequilibrium stationary states consider several
2980: thermo\-stats with interaction energy with $\CC_0$, $W_j(\V X_0,\V
2981: X_j)$, as in Eq. (\ref{9.1}). The equations of motion should be
2982: Eq. (\ref{18.2})
2983:
2984: In general solutions of Eq.(\ref{18.2}) {\it will not be quasi periodic} and
2985: the Chaotic Hypothesis, \cite{GC95b,Ga00,Ga07}, can be assumed: if so the
2986: dynamics should select an invariant distribution $\m$. The
2987: distribution $\m$ will give the statistical properties of the
2988: stationary states reached starting the motion in a thermostat
2989: configuration $(\V X_j,\V{{\dot X}}_j)_{j>0}$, randomly chosen with
2990: ``uniform distribution'' $\n$ on the spheres $m\V{{\dot X}}_j^2=3N_jk_B
2991: T_j$ and in a random eigenstate of $H$. The distribution $\m$, if
2992: existing and unique, could be named the {\it SRB distribution}
2993: corresponding to the chaotic motions of Eq.(\ref{18.2}).
2994:
2995: In the case of a system {\it interacting with a single thermostat} the
2996: latter distribution should be equivalent to the canonical
2997: distribution. As in Sec.\ref{sec11} an important consistency check for
2998: the model just proposed in Eq.(\ref{18.2}) is that there should exist
2999: at least one stationary distribution $\m$ equivalent to the canonical
3000: distribution at the appropriate temperature $T_1$ associated with the
3001: (constant) kinetic energy of the thermostat: $K_1=\frac32 k_B
3002: T_1\,N_1$. However also in this case, as already in Sec.\ref{sec11}, it
3003: does not seem possible to define a simple invariant distribution, not
3004: even in the adiabatic approximation. As in Sec.\ref{sec11}, equivalence
3005: between $\m$ and a Gibbs distribution at temperature $T_1$ can only be
3006: conjectured.
3007:
3008: Furthermore it is not clear how to define phase space contraction,
3009: hence how to formulate a FT, although the equations are reversible.
3010: %\vfill\eject
3011:
3012: \def\SEC{References}\iniz
3013: \label{secRef}
3014: {\nota\baselineskip=9pt
3015: \ifnum\biblio=0\bibliography{0Bibcaos}\fi
3016: \ifnum\biblio=1\input Ge.bbl \fi
3017: \bibliographystyle{unsrt}
3018: %\input Ge.bbl
3019: }
3020: %\vfill\eject
3021: %
3022: %$$ \Ba = \Bb = \Bg = \Bd = \Be = \Bee = \Bz = \Bh = \Bthh = \Bth = \Bi
3023: %= \Bk = \Bl = \Bm = \Bn =
3024: %$$
3025: %$$\Bx = \Bom = \Bp = \Br = \Bro = \Bs = \Bsi = \Bt = \Bu = \Bf = \Bff
3026: %= \Bch = \Bps =
3027: %$$
3028: %$$\Bo = \Bome = \BG = \BD = \BTh = \BL = \BX = \BP = \BS = \BU = \BF =
3029: %\BPs = \BO = \BDpr = \Bstl $$
3030: \end{document}
3031: