1:
2: In this chapter we address the issue of improving the anneal bound
3: on the critical line~$h_c(\cdot)$ of the random copolymer
4: via the so--called constrained annealing, that means nothing but
5: applying the annealing procedure (which is just Jensen's inequality)
6: after having added to the Hamiltonian a disorder--dependent term
7: (sometimes interpreted as a Lagrange multiplier) in a way that
8: the quenched expressions are left unchanged, remember \S~\ref{sec:ph_random}
9: of Chapter~\ref{ch:first}.
10:
11: A popular class of multipliers is the one consisting of empirical
12: averages of local functions of the disorder.
13: These multipliers are particularly suitable for computations,
14: and it is often believed that in this
15: class one can approximate arbitrarily well the quenched free energy.
16:
17: We are going to prove that {\sl this is not the case}
18: for a wide family of polymer models, including
19: the copolymer near a selective interface and the pinning/wetting models
20: defined in Chapter~\ref{ch:first}.
21: More precisely we show that the multipliers in the above class
22: cannot improve on the basic annealed bound from the viewpoint
23: of characterizing the phase diagram. For simplicity the proof has
24: been carried out under the assumption that the random variable~$\go_1$
25: takes only a finite number of values, however the statement remains true
26: also the general case, provided one makes some suitable boundedness
27: assumptions on the multiplier.
28:
29: \smallskip
30: The article~\cite{cf:CG} has been taken from the content of this chapter.
31:
32: \smallskip
33: \section{The framework and the main result}
34: \label{sec:intro_CG}
35:
36:
37: \smallskip
38: \subsection{The general set--up}
39: A number of disordered models of linear chains undergoing
40: localization or pinning effects can be put into the following general framework.
41: Let $S:=\left\{ S_n \right\}_{n=0,1,\ldots}$ be a process with $S_n$ taking
42: values in $\Z^d$, $d \in \N :=\{1,2,\ldots\}$ and
43: law $\bP$.
44:
45:
46:
47: The disorder in the system is given by a sequence $\go :=\left\{\go _n\right\}_n$
48: of IID random variables taking values in a finite set $\Gamma$ with law $\bbP$, acting on the path of $S$ via an Hamiltonian
49: that, for a system of size $N$, is a function $H_{N,\go}$
50: of the trajectory $S$, but depending only on $S_0, S_1, \ldots, S_N$.
51: One is interested in the properties of the probability measures
52: $\bP_{N, \go}$ defined by giving the density with respect to $\bP$:
53: \begin{equation}
54: \label{eq:1}
55: \frac{\dd \bP_{N, \go}}{\dd\bP} \left( S \right) \, =\,
56: \frac1{Z_{N, \go}} \exp\left(H_{N, \go}\left(S\right)\right),
57: \end{equation}
58: where $Z_{N, \go}:= \bE \left[\exp\left(H_{N, \go}\left(S\right)\right)\right]$ is the normalization constant.
59: Our attention focuses on the
60: asymptotic behavior of $\log Z_{N, \go}$.
61:
62: In the sequel we will assume:
63:
64: \medskip
65: \begin{assumption} \rm
66: There exists a sequence $\left\{ D_n \right\}_n$
67: of subsets of $\Z^d$ such that
68: $\bP ( S_n \in D_n \text{ for } n=1,2, \ldots, N)\stackrel{N\to \infty}{\asymp} 1 $, namely
69: \begin{equation}
70: \lim_{N\to \infty} \frac 1N \log \bP \left( S_n \in D_n \text{ for } n=1,2, \ldots, N\right)=0,
71: \end{equation}
72: and $H_{N, \go}(S)=0$ if $S_n \in D_n$ for $n=1, 2, \ldots, N$.
73: \end{assumption}
74:
75: \vskip 0.1 cm
76:
77: One sees directly that this hypothesis implies
78: \begin{equation}
79: \label{eq:first}
80: \liminf_{N \to \infty}
81: \frac 1N \log Z_{N, \go} \, \ge
82: \lim_{N \to \infty}
83: \frac 1N \log \bP \left( S_n \in D_n \text{ for } n=1,2, \ldots, N\right)\, =\, 0,
84: \end{equation}
85: $\bbP (\dd \go)$--a.s.. We will assume that $\left\{(1/N)\log Z_{N, \go}\right\}_N$
86: is a sequence of integrable random variables that converges
87: in the $L^1\left( \bbP (\dd \go)\right)$ sense and
88: $P(\dd \go)$--almost surely to a constant, the free energy, that we will call $\free$.
89: These assumptions are verified in the large majority of the interesting situations,
90: for example whenever super/sub--additivity tools are applicable.
91:
92:
93: Of course \eqref{eq:first} says that $\free \ge 0$ and one is lead to
94: the natural question of whether $\free =0$ or $\free >0$.
95: In the instances that we are going to consider the free energy may be
96: zero or positive according to some parameters from which the $H_{N, \go}$
97: depends: $\free =0$ and $\free >0$ are associated to sharply different
98: behaviors of the system.
99:
100: \medskip
101:
102: In order to establish upper bounds on $\free$
103: one may apply directly Jensen inequality ({\sl annealed bound}) obtaining
104: \begin{equation}
105: \label{eq:J}
106: \free \, \le \, \liminf_{N \to \infty} \frac 1N \log \bbE \left[ Z_{N,\go}
107: \right]\, =: \, \freea,
108: \end{equation}
109: and, in our context,
110: if $\freea=0$ then $\free =0$.
111: The annealed bound
112: may be improved by adding to
113: $H_{N, \go}(S)$ an integrable function $A_N: \Gamma^\N \to \R$ such that
114: $\bbE\left[ A_N (\go) \right]=0$: while the left--hand side is unchanged,
115: $\freea$ may depend on the choice of $\{A_N\}_N$.
116: We stress that not only $\free$ is left
117: unchanged by $H_{N, \go}(S) \to H_{N, \go}(S)+A_N (\go)$, but $\bP_{N,\go}$
118: itself is left unchanged (for every $N$).
119: Notice that the choice
120: $A_N (\go) = -\log Z_{N,\go} + \bbE\left[ \log Z_{N,\go}\right]$
121: yields the equality in \eqref{eq:J}.
122:
123: In the sequel when we refer to $\freea$ we mean that $Z_{N,\go}$ is defined with respect to $H_{N, \go}$
124: satisfying the Basic Hypothesis (no $A_N$ term added).
125:
126: \smallskip
127: \subsection{The result}
128: What we prove in this note is that
129: \medskip
130:
131: \begin{proposition}
132: \label{th:main}
133: If $\freea >0$ then
134: for every local function $F: \Gamma ^\N \longrightarrow \R$ such that $\bbE \left[ F (\go) \right]=0$
135: one has
136: \begin{equation}
137: \label{eq:main_CG}
138: \liminf_{N \to \infty} \frac 1 N \log
139: \bbE \bE \left[ \exp \left( H_{N, \go}(S) + \sum_{n=0}^N F(\theta_n \go)\right)\right]\, > \, 0,
140: \end{equation}
141: where $(\theta_n \go)_m= \go_{n+m}$.
142: \end{proposition}
143:
144: \medskip
145:
146: We can sum up this result by saying that when $ \free =0$ but $ \freea >0$
147: it is of no use modifying the Hamiltonian by adding the
148: empirical average of a (centered) local function.
149:
150: On a mathematical level it is clear that we are playing with
151: an exchange of limits and that it is not obvious that the
152: free energy, recall the optimal choice of $A_N$ above,
153: may be approximated via empirical averages of a
154: local function of the disorder. But we remark
155: that in the physical literature the approach of approximating
156: the free energy via what can be viewed
157: as a constrained annealed computation, the term $\sum_{n=0}^N F(\theta_n \go)$
158: being interpreted as a Lagrange multiplier,
159: is often considered as an effective way of approximating the quenched
160: free energy. Here we mention in particular \cite{cf:Morita}
161: and \cite{cf:Kuhn} in which this approach is taken up in a systematic way:
162: the aim is to approach the quenched free energy by constrained
163: annealing via local functions $F$ that are more and more complex,
164: the most natural example being linear combinations of correlations of higher and higher
165: order.
166: \smallskip
167:
168: The proof of
169: Proposition \ref{th:main} is based on the simple observation
170: that whenever $A_N$ is centered
171: \begin{multline}
172: \label{eq:argproof}
173: \frac 1 N \log
174: \bbE \bE\left[ \exp \left( H_{N, \go}(S) + A_N(\go) \right)\right]\, \ge
175: \\
176: \frac 1 N \log
177: \bbE \left[ \exp \left( A_N(\go) \right)\right] +
178: \frac 1 N \log \bP \left( S_n \in D_n \text{ for } n=1, 2, \ldots , N
179: \right)\, =:\, Q_N+P_N.
180: \end{multline}
181: By hypothesis $P_N=o(1)$ so one has to consider the asymptotic behavior
182: of $Q_N$. If $\liminf_N Q_N >0$ there is nothing to prove.
183: So let us assume that $\liminf_{N}Q_N=0$:
184: in this case the inferior limit of the left--hand side of \eqref{eq:argproof}
185: may be zero and we want to exclude this possibility
186: when $\tilde f >0$ and $A_N (\go)= \sum_{n=0}^N F(\theta_n \go)$, $F$ local and centered
187: (of course in this case $\lim_N Q_N$ does exist).
188: And in Theorem~\ref{th:main1} below in fact
189: we show that
190: if $\log \bbE\left[ \exp\left(A_N(\go)\right)\right] =o(N)$, then
191: there exists a local function $G$ such that $F( \go)= G(\theta_1 \go)- G(\go)$
192: so that $ \{\sum_{n=0}^N F(\theta_n \go)\}_N$ is just a boundary term and
193: the corresponding constrained annealing is just the standard annealing.
194:
195: Notice that having chosen $\Gamma$ finite frees us from
196: integrability conditions.
197:
198:
199:
200: \medskip
201:
202: \begin{rem}
203: \label{th:rem}
204: \rm
205: We stress that our Basic Hypothesis is more general than it may look at first.
206: As already observed, one has the freedom of adding to the Hamiltonian $H_{N, \go}(S)$
207: any term that does not depend on $S$ (but possibly does depend on $\go$ and $N$)
208: without changing the model $ \bP_{N,\go}$.
209: It may therefore happen
210: that the {\sl natural} formulation of the Hamiltonian does not satisfy
211: our Basic Hypothesis, but it does after a suitable additive correction.
212: This happens for example for the Copolymer near a selective interface model,
213: as we have seen in~\S~\ref{rem:Z} of Chapter~\ref{ch:first} (see also
214: \S~\ref{sec:cop} below):
215: the additive correction in this case is linear in $\go$ and it corresponds to
216: what in \cite{cf:ORW} is called {\sl first order} Morita approximation.
217: In these terms, Proposition \ref{th:main} is saying that
218: {\sl higher order} Morita approximations cannot improve
219: the bound on the critical curve found with the first order computation.
220: \end{rem}
221:
222: \smallskip
223:
224: Let us now look at applications of Proposition \ref{th:main}.
225:
226: \smallskip
227: \subsection{Random rewards or penalties at the origin}
228: Let $S$, $S_0=0\in \Z ^d$, be a random walk with centered IID non degenerate increments $\{X_n\}_n$,
229: $(X_n)_j\in \{-1,0,1\}$ for $j=1, 2, \ldots, d$,
230: and
231: \begin{equation}
232: H_{N, \go} = \gb
233: \sum_{n=1}^N \left(1+ \gep \go_n\right) \ind_{\{S_n=0\}}.
234: \end{equation}
235: for $\gb \ge 0$ and $\gep \ge 0$. This model is a $d$--dimensional
236: version (with somewhat different notations) of
237: the pinning model introduced in~\S~\ref{sec:pinning} of Chapter~\ref{ch:first}.
238: The random variable $\go_1$ is chosen
239: such that $\bbE[\exp (\gl\go_1)]<\infty$ for every $\gl\in\R$, and centered.
240: We write $\free (\gb , \gep)$ for $\free$:
241: by super--additive arguments $\free$ exists and it is self--averaging (this
242: observation is valid for all the models we consider and will not be repeated).
243: As we already remarked in Chapter~\ref{ch:first}, for $\gep =0$ the model can be solved, see
244: e.g. \cite{cf:G}, and in particular $\free (\gb, 0)=0$ if and only if $\gb \le \gb_c (d):= -\log (1-\bP(S$ never
245: comes back to $0))$. Adding the disorder makes this
246: model much more complex: the annealed bound yields $\free (\gb ,\gep) =0$
247: if $ \gb \le \gb_c(d)- \log \bbE \left[\exp (\gep \go_1)\right]
248: =:\tilde {\gb_c}$.
249: It is an open question whether $\tilde {\gb_c}$ coincides with
250: the quenched critical value or not, that is whether $\free (\gb , \gep) =0$
251: implies $ \gb \le \tilde {\gb_c}$ or not. For references about this issue
252: we refer to \cite{cf:AS,cf:Pet},
253: see however also the next
254: paragraph: the model we are considering can in fact be exactly mapped
255: to the wetting problem (\cite{cf:AS}, \cite{cf:G}).
256: Proposition~\ref{th:main} applies to this context with $D_n= \{0\}^\complement$ for every $n$
257: \cite[Ch. 3]{cf:Feller}
258: and says that one cannot answer this question via constrained annealed bounds.
259:
260:
261:
262: \smallskip
263: \subsection{Wetting models in $1+d$ dimensions}
264: Let $S$ and $\go$ as in the previous example
265: and
266: \begin{equation}
267: \label{eq:Wetting}
268: H_{N, \go} =
269: \begin{cases}
270: \gb \sum_{n=1}^N \left( 1 + \gep \go_n\right) \ind_{\{(S_n)_d=0\}}
271: &\text{ it } (S_n)_d\ge 0 \text{ for } n=1,2, \ldots , N \\
272: -\infty &\text{ otherwise.}
273: \end{cases}
274: \end{equation}
275: with $\gb \ge 0$ and
276: $\gep\ge 0$.
277: If one takes the directed walk viewpoint,
278: that is if one considers the walk $\{(n,S_n)\}_n$,
279: then this is a model of a walk constrained above
280: the (hyper--)plane $x_d=0$ and rewarded $\gb$, on the average,
281: when touching this plane. If $d=1$ then this is an effective
282: model for a (1+1)--dimensional interface above a wall which mostly attracts it.
283: As a matter of fact in this case there is no loss of generality
284: in considering $d=1$, since in the directions parallel to the wall
285: the model is just the original walk.
286: Once again if $\gep =0$ the model can be solved in detail,
287: see e.g. \cite{cf:G}. Computing the critical $\gb$
288: and deciding whether the annealed bound is sharp, at least
289: for small $\gep$, is an unresolved and disputed question
290: in the physical literature, see e.g. \cite{cf:FLNO}, \cite{cf:DHV} and \cite{cf:TC}.
291: Proposition \ref{th:main} applies with the choice
292: $D_n =\bbZ ^{d-1}\times \N$.
293:
294:
295:
296: \smallskip
297: \subsection{Copolymer and adsorption models}
298: \label{sec:cop}
299: Choose $S$ as above and take the directed walk viewpoint.
300: Imagine that above the axis ($x_d>0$) is filled of a solvent $A$,
301: while below ($x_d<0$) there is a solvent $B$. At $x_d=0$ there is the
302: interface. We choose $\go = \{A,B\}$ and for example
303: \begin{equation} \label{eq:hab}
304: H^{AB}_{N, \go}(S)\, =\,
305: \sum_{n=1}^N \left(a \ind_{\{\sign(S_n)=+1,\, \go_n =A\}} +
306: b\ind_{\{\sign (S_n) =-1, \, \go_n =B\}}+ c\ind_{\{S_n=0\}}
307: \right)
308: \end{equation}
309: with $a$, $b$ and $c$ real parameters, $\sign (S_n) := \sign \big( (S_n)_d \big)$
310: and the convention $\sign(S_n)= \sign(S_{n-1})$ if $(S_n)_d=0$
311: already used in Chapter~\ref{ch:first}.
312: In order to apply Proposition \ref{th:main}
313: one has to subtract a disorder dependent term, cf. Remark~\ref{th:rem}: if $a \ge b$
314: we change the Hamiltonian
315: \begin{equation}
316: \label{eq:H1}
317: H_{N, \go}(S)\, :=\, H^{AB}_{N, \go}(S)- \sum_{n=1}^N a \ind_{\{ \go_n =A\}}.
318: \end{equation}
319: without changing the measure $\bP_{N, \go}$ while the free energy
320: has the trivial shift
321: from $\free$ to $\free - a \bbP \left( \go_1=A\right)$.
322: One can therefore choose $D_n =\bbZ ^{d-1}\times \N$ and Proposition \ref{th:main}
323: applies. This model has been considered for example
324: in \cite{cf:ORW}.
325:
326: Note that if~$c=0$ and~$d=1$ the model is nothing but the copolymer model introduced
327: in Chapter~\ref{ch:first}, that is we can cast~\eqref{eq:hab} in the form
328: \begin{equation}
329: \label{eq:H2}
330: H_{N,\go} (S)\, = \, \gl \sum_{n=1}^N \left(\go_n +h \right) \sign (S_n) \,,
331: \end{equation}
332: with $\go$ taking values in $\R$. Once again the Hamiltonian
333: has to be corrected by subtracting
334: the term $\gl \sum_n (\go_n +h)$ (which is exactly what was done in \S~\ref{rem:Z} of
335: Chapter~\ref{ch:first}) in order to apply Proposition \ref{th:main}.
336: One readily sees that \eqref{eq:H1}
337: and \eqref{eq:H2} are the same model when in the second case
338: $\go$ takes only the values $\pm 1$, $A=+1$ and $B=-1$,
339: and $h= (a-b)/(a+b)$, $\gl= (a+b)/4$.
340:
341:
342: Proposition \ref{th:main} acquires some interest in this context:
343: in fact we have already remarked that
344: the physical literature is rather split on the
345: precise value of the critical curve and on whether the annealed
346: bound is sharp or not.
347: We recall that the numerical analysis performed in Chapter~\ref{ch:cgg}
348: is suggesting that the annealed curve does not coincide with
349: the quenched one, and in view of Proposition~\ref{th:main} this would mean that
350: constrained annealing via local functions cannot capture
351: the phase diagram of the quenched system.
352:
353:
354:
355:
356: \smallskip
357: \subsection{Further models and observations}
358: In spite of substantial numerical evidence
359: that in several instances $\free =0$ but $\freea>0$,
360: we are unaware of an {\sl interesting} model for which
361: this situation is rigorously known to happen.
362: Consider however the case $\bbP (\go_n =\pm 1)=1/2$ and
363: \begin{equation}
364: H_{N, \go} (S)=
365: \gb \sum_{n=1}^N \left( 1 + \gep \go_n\right) \ind_{\{S_n=n\}},
366: \end{equation}
367: with $\gb$ and $\gep$ real numbers and $S$ the simple random walk on $\Z$.
368: We observe that
369: Proposition \ref{th:main} applies to this case with $D_n =\{ n\}^\complement$ and
370: that the model is solvable in detail.
371: In particular $\free (\gb, \gep) = (\gb - \log 2)\vee 0$, regardless of the value of $\gep$.
372: The annealed computation instead yields $\freea (\gb , \gep)=
373: (\gb +\log \cosh(\gep)-\log 2)\vee 0$. Notice in particular
374: that the critical values of $\gb$, respectively $\log 2$ and
375: $\log 2 -\log \cosh(\gep)$, differ as long as there is disorder
376: in the system ($\gep \neq 0$).
377: It is interesting to see in this toy model how
378: $A_N$ has to be chosen {\sl very non local} in order to
379: improve on the annealed bound.
380:
381: \smallskip
382:
383: \begin{rem}\rm
384: We point out that we restricted our examples
385: only to cases in which $S$ is a simple random walk, but
386: in principle our approach goes through for much more general models,
387: like walks with correlated increments or self--interacting walks, see
388: \cite{cf:OTW} for an example.
389: And of course $S_n$ takes values in $\Z^d$ only for ease of exposition
390: and can be easily generalized.
391: It is however unclear whether our argument applies
392: to the disordered wetting problem in $d+1$ dimensions, $d>1$.
393: In this case $S$ is a random interface, the Hamiltonian is like
394: in
395: \eqref {eq:Wetting}, but $n \in \{0,1, 2, \ldots\} ^d$, $S_n\in \Z$ or
396: $\R$. We set for example $S_n=0$ when one of the coordinates
397: of $n$ is zero. The missing ingredient is an analog of Theorem
398: \ref{th:main1} in higher dimensions.
399: \end{rem}
400:
401:
402:
403: \smallskip
404: \section{On cocycles with null free energy}
405: \label{sec:cocycles}
406:
407: Let $\{\bs{\go}_n\}_{n\in\N}$ be an IID sequence of random variables under the probability measure $\bbP$, taking values in a finite space $\Gamma$ (we have switched the notation $\go \to \bs{\go}$ for clarity). The law of $\bs{\go}_1$ on $\Gamma$ is denoted by $\nu$: we will assume that $\nu(\ga)>0$ for all $\ga\in\Gamma$.
408:
409: We are interested in families $A = \{A_N\}_{N\in\N}$ of random variables of the form of empirical averages of a centered local function $F$, that is
410: \begin{equation} \label{eq:hamiltonian}
411: A_N = \sum_{n = 1}^N F(\bs{\go}_n,\ldots,\bs{\go}_{n+k})\,,
412: \end{equation}
413: where $k\in\{0\} \cup \N$ and $F$ is a real function defined on $\Gamma ^{k+1}$ such that $\int F \dd\nu^{*(k+1)}=0$. We will call $A = \{A_N\}_{N\in\N}$ a centered \textsl{cocycle}, and with some abuse of notation we will speak of {\sl the cocycle~$F$} to mean the cocycle~$\{A_N\}_{N\in\N}$ defined by ~(\ref{eq:hamiltonian}).
414:
415: A cocycle $F:\Gamma^{k+1} \to \R$ is said to be a \textsl{coboundary} if (when $k\geq 1$) there exists a function $G:\Gamma ^k \to \R$ such that
416: \begin{equation} \label{eq:grad}
417: F(\ga_1,\ldots ,\ga_{k+1}) = G(\ga_2,\ldots \ga_{k+1}) - G(\ga_1,\ldots ,\ga_{k})
418: \end{equation}
419: for all $\ga_1,\ldots, \ga_{k+1} \in \Gamma$. When $k=0$, we say that $F$ is a \textsl{coboundary} if it is identically zero: $F(\ga) = 0$ for every $\ga \in \Gamma$.
420:
421: \smallskip
422:
423: For $\gb \in\R$ we define the free energy $L^F(\gb)$ of a cocycle $F$ as
424: \begin{equation} \label{eq:free_energy}
425: L^F(\gb) := \lim_{N\to\infty} \frac 1N \log \bbE \Big[e^{\gb A_N}\Big] \,.
426: \end{equation}
427: The limit above is easily seen to exist by a standard superadditive argument, and Jensen's inequality yields immediately $L^F(\gb)\geq0$. Of course, if $F$ is a coboundary then the corresponding free energy vanishes for all $\gb\in\R$. That also the converse is true is the object of the following theorem.
428:
429: \begin{theorem} \label{th:main1}
430: Let $F$ be a centered cocycle, and let $L^F(\gb)$ be the corresponding free energy, defined by (\ref{eq:free_energy}). The following conditions are equivalent:
431: \begin{enumerate}
432: \item $F$ is a coboundary;
433: \item $L^F(\gb)=0$ for all $\gb\in\R$;
434: \item $L^F(\gb_0)=0$ for some $\gb_0\in\R\setminus\{0\}$.
435: \end{enumerate}
436: \end{theorem}
437:
438: The proof is obtained combining convexity ideas with the following combinatorial reformulation of the condition that a function be a coboundary.
439:
440: \begin{lemma} \label{lem:main}
441: A function $F:\Gamma ^{k+1} \to \R$ is a coboundary if and only if for every $N\in\N$ and for every $(\eta_1,\ldots , \eta_{N}) \in \Gamma^N$ the following relation holds:
442: \begin{align} \label{eq:grad1}
443: \sum_{i=1}^{N} F(\eta_i,\eta_{i\oplus_N 1},\ldots, \eta_{i\oplus_N k}) = 0 \,,
444: \end{align}
445: where for $a,b \in \N$ we have set $a\oplus_N b := (a + b)\mod N$.
446: \end{lemma}
447:
448: \proof
449: The {\sl if} part trivially follows from the definition of a coboundary (see (\ref{eq:grad})), so we can focus on the
450: {\sl only if} part. As a matter of fact, we will use the hypothesis of the Lemma only for two values of $N$, namely $N=2k$ and $N=2k+1$.
451:
452: Let us take $k$ elements $\gamma_1,\ldots,\gamma_k\in \Gamma$,
453: arbitrarily chosen, that will be kept \textsl{fixed} throughout the proof; moreover, let $\ga_1,\ldots , \ga_{k+1}$ denote generic elements of~$\Gamma$. We start rewriting equation~(\ref{eq:grad1}) for $N=2k+1$, with $(\eta_1,\ldots , \eta_{N})=(\ga_1,\ldots , \ga_{k+1},\gamma_1,\ldots,\gamma_k)$, as
454: \begin{align} \label{eq:first_step}
455: F(\ga_1,\ldots , \ga_{k+1}) = -\sum_{i=1}^{k} F(\ga_{i+1},\ldots, \ga_{k+1}, \gamma_1,\ldots, \gamma_i) - \sum_{i=1}^{k} F(\gamma_i,\ldots, \gamma_k,\ga_1,\ldots, \ga_i) \,.
456: \end{align}
457: In order to determine an alternative expression for the second sum in the r.h.s., we use again equation~(\ref{eq:grad1}), this time with~$N=2k$ and~$(\eta_1,\ldots , \eta_{N})=(\ga_1,\ldots , \ga_{k},\gamma_1,\ldots,\gamma_k)$, getting
458: \begin{equation} \label{eq:second_step}
459: \sum_{i=1}^{k} F(\gamma_i,\ldots, \gamma_k,\ga_1,\ldots, \ga_i) = - \sum_{i=1}^{k} F(\ga_i,\ldots, \ga_k,\gamma_1,\ldots, \gamma_i)\,.
460: \end{equation}
461:
462: If now we introduce a function $G:\Gamma^k \to \R$, defined by
463: \begin{align*}
464: G(\zeta_1,\ldots, \zeta_k) := - \sum_{i=1}^{k} F(\zeta_i,\ldots, \zeta_k,\gamma_1,\ldots, \gamma_i) \,,
465: \end{align*}
466: we can combine equations (\ref{eq:first_step}) and (\ref{eq:second_step}) to get
467: \begin{align*}
468: F(\ga_1,\ldots , \ga_{k+1})
469: %& = -\sum_{i=1}^{k} F(\ga_{i+1},\ldots, \ga_{k+1}, \gamma_1,\ldots, \gamma_i) + \sum_{i=1}^{k} F(\ga_i,\ldots, \ga_k,\gamma_1,\ldots, \gamma_i) \nonumber \\
470: = G(\ga_2,\ldots \ga_{k+1}) - G(\ga_1,\ldots \ga_{k}) \,,
471: \end{align*}
472: so that the proof is completed.
473: \qed
474:
475: \bigskip
476:
477: \noindent\textbf{Proof of Theorem \ref{th:main1}.} It has already been remarked that $(1) \Rightarrow (2)$, and of course $(2) \Rightarrow (3)$ holds trivially. In the following we are going to prove that $(3) \Rightarrow (2) \Rightarrow (1)$.
478:
479: We start determining an explicit expression for the free energy. For this, we define a slight modification of the cocycle $A$ defined by~(\ref{eq:hamiltonian}), by setting
480: \begin{equation} \label{eq:mod_ham}
481: \tilde{A}_N := \sum_{n = 1}^N F(\bs{\go}_n,\bs{\go}_{n\oplus_N 1},\ldots,\bs{\go}_{n\oplus_N k})\,,
482: \end{equation}
483: where by $\oplus_N$ we mean addition modulo $N$. Of course, only the last $k$ addends in the sum are really changed: as $F$ is a bounded function (the space $\Gamma$ is finite), it easily follows that the free energies of $A$ and $\tilde{A}$ are the same, so that we can write
484: \begin{equation} \label{eq:zeta}
485: L^F(\gb) = \lim_{N\to\infty} \frac 1 N \log Z_N(\gb) \qquad \text{where} \qquad Z_N(\gb) = Z^F_N(\gb) = \bbE \Big[ e^{\gb \tilde{A}_N} \Big] \,.
486: \end{equation}
487:
488: Now we introduce the $\Gamma^{k+1} \times \Gamma^{k+1}$ matrix $A_\gb$, defined for $\ga_i,\gamma_i \in \Gamma,\ i=1,\ldots,k+1$ by
489: \begin{equation} \label{eq:matrix}
490: A_\gb \big[ (\ga_1, \ldots, \ga_{k+1}), (\gamma_1, \ldots, \gamma_{k+1}) \big] := \gd_{\gamma_1,\ga_2} \cdots \gd_{\gamma_{k}, \ga_{k+1}} \cdot e^{\gb F(\gamma_1, \ldots, \gamma_{k+1})} \cdot \nu(\gamma_{k+1})\,.
491: \end{equation}
492: Developing the expectation defining $Z_N(\gb)$ we get
493: \begin{align}
494: Z_N(\gb) &\;=\; \sum_{\zeta_1,\ldots,\zeta_N \in \Gamma} e^{\gb \sum_{i=1}^N F(\zeta_i,\zeta_{i\oplus_N 1}, \ldots, \zeta_{i\oplus_N k})} \cdot \nu(\zeta_1) \cdots \nu(\zeta_N) \nonumber \\
495: \label{eq:trace}
496: & \;=\; \text{Tr} \big[ A_\gb ^N \big] \;=\; \sum_{i=1}^{|\Gamma|^{2(k+1)}} e_i(\gb)^N \,,
497: \end{align}
498: where $\{e_i(\gb),\ i=1,\ldots,|\Gamma|^{2(k+1)}\}$ are the (possibly complexes) eigenvalues of the matrix $A_\gb$ (counted repeatedly according to their algebraic multiplicity). It's immediate to check that $A_\gb$ is an irreducible, aperiodic matrix, and since its entries are nonnegative we can apply Perron--Frobenius theory \cite{cf:Asm}: there exists a real positive simple eigenvalue, say $e_1(\gb)$, such that $|e_i(\gb)| < e_1(\gb)$ for every $i\geq 2$. To lighten the notation, from now on we will let $e(\gb) := e_1(\gb)$. Combining (\ref{eq:zeta}) with (\ref{eq:trace}) we get
499: \begin{equation} \label{eq:zeta_as}
500: Z_N(\gb) = e(\gb)^N \cdot \Bigg( 1 + \sum_{i=2}^{|\Gamma|^{2(k+1)}} \bigg( \frac{e_i(\gb)}{e(\gb)} \bigg) ^N \Bigg) \,,
501: \end{equation}
502: so that
503: \[
504: Z_N(\gb) \cdot e(\gb)^{-N} \to 1 \quad \text{as }N\to\infty \,.
505: \]
506: From this \textsl{sharp asymptotics} for $Z_N(\gb)$ we obtain in particular the explicit expression of $L^F(\gb)$ we were looking for:
507: \begin{equation} \label{eq:log_e}
508: L^F(\gb) = \log e(\gb) \,.
509: \end{equation}
510:
511: This equation shows that $L^F(\gb)$ is a \textsl{real analytic} function of $\gb\in\R$, since $e(\gb)$ is so: this is because the Perron--Frobenius eigenvalue is a simple root of the characteristic polynomial and the entries of $A_\gb$ are real--analytic functions of $\gb\in\R$.
512:
513: From (\ref{eq:zeta}) it is clear that $\log Z_N(\gb)$ is a convex function of $\gb\in\R$, for every $N\in\N$. Moreover, we have $Z_N(\gb) \geq 1$ for every $\gb\in\R$ by Jensen's inequality, and trivially $Z_N(0)=1$. It follows immediately that $L^F(\gb)$ is a convex function too, being the pointwise limit of $\log Z_N(\gb)/N$, that $L^F(\gb) \geq 0$ for every $\gb\in\R$, and $L^F(0)=0$.
514:
515: \smallskip
516:
517: Let's assume that condition $(3)$ in the statement of the theorem holds, that is $L^F(\gb_0)=0$ for some $\gb_0 > 0$ (the case $\gb_0<0$ is completely analogous): the preceding observations yield $L^F(\gb)=0$ for every $\gb \in [0,\gb_0]$, and by analyticity we conclude that indeed $L^F(\gb)=0$ for every $\gb \in \R$. We have thus shown that $(3) \Rightarrow (2)$.
518:
519: \smallskip
520:
521: Now we assume that condition $(2)$ holds: by (\ref{eq:log_e}) this means $e(\gb)=1$ for every $\gb\in\R$, and (\ref{eq:zeta_as}) we have that
522: \begin{equation} \label{eq:ineq}
523: |Z_N(\gb)| \leq e(\gb)^N \cdot |\Gamma|^{2(k+1)} = |\Gamma|^{2(k+1)} \qquad \forall N\in\N\,,\ \forall \gb\in\R\,.
524: \end{equation}
525: Since $\log Z_N(\gb)$ is a convex function, $Z_N(\gb)$ is convex too; furthermore, we have already remarked that $Z_N(\gb) \geq 1$ for every $\gb\in\R$ and that $Z_N(0)=1$. Since (\ref{eq:ineq}) shows that $|Z_N(\gb)|$ is bounded, by elementary convex analysis it follows that $Z_N$ must be constant, therefore $Z_N(\gb)=1$ for all $\gb\in\R$ and $N\in\N$. This means that for every $\gb\in\R$ Jensen's inequality for $Z_N(\gb)$ it's not strict: since for any $\gb>0$ the function $\{x \mapsto e^{\gb x}\}$ is a strictly convex function, this can happen if and only if $\tilde{A}_N$ is $\bbP$--a.s. constant, for every $N\in\N$. Recalling (\ref{eq:mod_ham}) and the fact that by hypothesis $\nu(\ga)>0$ for every $\ga\in\Gamma$, this amounts to saying that
526: \[
527: \sum_{i=1}^{N} F(\eta_i,\eta_{i\oplus_N 1},\ldots, \eta_{i\oplus_N k}) = 0 \,,
528: \]
529: for every $N\in\N$ and for every $\eta_1, \ldots, \eta_N \in \Gamma$: applying Lemma~\ref{lem:main} we conclude that $F$ is a coboundary, and the proof is complete.\qed
530:
531: