math0511561/cgz.tex
1: 
2: 
3: In this chapter we consider a general model
4: of an heterogeneous polymer chain in the proximity of
5: an interface between two selective solvents, which
6: includes as special cases the copolymer near a selective interface
7: and the pinning model introduced in Chapter~\ref{ch:first}.
8: The heterogeneous character of the model comes from the fact that
9: the interaction of each {\sl monomer unit} is governed by a
10: {\it charge} that it carries. We consider the model in the {\sl periodic setting},
11: that is when the charges repeat themselves along the chain in a periodic fashion.
12: The main question is of course whether the polymer remains tightly close
13: to the interface ({\sl localization})
14: or there is a marked preference for one solvent ({\sl delocalization}).
15: 
16: We propose an approach based on renewal theory
17: that yields sharp estimates on the partition function of the model
18: in all the regimes (localized, delocalized and critical).
19: This in turn allows to get a very precise
20: description of the polymer measure, both in a local sense
21: ({\sl thermodynamic limit}) and in a global sense ({\sl scaling limits}):
22: see~\S~\ref{sec:spirit} for an outline of our results
23: and~\S~\ref{sec:results} for a detailed exposition.
24: A key point, but also a byproduct, of our analysis is the closeness of
25: the polymer measure to a suitable {\sl Markov Renewal Process}.
26: 
27: \smallskip
28: The preprint~\cite{cf:CGZ} has been taken from the content of this chapter.
29: 
30: \smallskip
31: \section{Introduction and main results}
32: \label{sec:intro_cgz}
33: 
34: \smallskip
35: 
36: \subsection{Two motivating models}
37: \label{sec:2examples}
38: We slightly enlarge our setting with respect to Chapter~\ref{ch:first}, namely
39: we work with a random walk $S:=\left\{ S_n\right\}_{n=0,1, \ldots}$
40: with IID symmetric increments $\{X_j\}_{j \ge 1}$ taking values in $\{-1, 0,+1\}$.
41: Hence the law of the walk is identified by
42: $p:=\bP \left( X_1 =1\right)(=\bP \left( X_1 =-1\right))$, and we assume that $p\in (0,1/2)$.
43: Note that we have excluded the case $p=1/2$ and this has been done in order to
44: lighten the exposition: all the results we  present have a close analog in the case $p=1/2$,
45: however the statements
46: require a minimum of notational care because of the periodicity of the walk. We also consider
47: a sequence $\go:=\left\{ \go_n\right\}_{n\in \N=\{1,2,\ldots\} }$ of real numbers
48: with the property that $\go_n=\go_{n+T}$ for some $T\in \N$ and for every $n$:
49: we denote by  $T(\go)$ the minimal value of $T$.
50: 
51: \smallskip
52: 
53: Before defining the general model that will be the object of our analysis,
54: we recall the two motivating models that were introduced in Chapter~\ref{ch:first}.
55: 
56:  \smallskip
57:  \begin{enumerate}
58:  \item {\it Pinning and wetting models.} For $\gl \ge 0$ consider the probability measure $\bP_{N,\go}$
59:   defined by
60:  \begin{equation}
61:  \label{eq:pinning}
62:  \frac{\dd \bP_{N, \go}}{\dd \bP} (S) \, \propto \, \exp \left( \gl \sum _{n=1}^ {N}
63:  \go_n \ind_{\left\{S_n=0 \right\}} \right).
64:  \end{equation}
65: The walk receives a {\sl pinning reward}, which may be negative or positive, each time
66: it visits the origin. By considering the directed walk viewpoint, that is $\left\{ (n, S_n)\right\}_n$,
67: one may interpret this model in terms of a directed linear chain
68: receiving an energetic contribution when it touches an interface. In this context it is natural
69: to introduce the asymmetry parameter $h:= \sum_{n=1}^T \go_n /T$, so that one isolates
70: a constant drift term from the {\sl fluctuating} behavior of $\go$.
71: The question is whether for large $N$
72: the measure $\bP_{N, \go}$ is rather attracted or repelled by the interface (there is in principle
73: the possibility for the walk to be essentially indifferent of such a change of measure, but we anticipate
74: that this happens only in trivially degenerate cases while in {\sl critical} situations a more subtle
75: scenario shows up).
76: 
77: By multiplying the right--hand side of  \eqref{eq:pinning} by $\ind_{\left\{S_n\ge 0:\,  n=1, \ldots, N\right\}}$
78: one gets to a so called {\sl wetting model}, that is the model of an interface interacting with an impenetrable
79: wall. The {\sl hard--wall} condition induces a repulsion effect of purely entropic origin which
80: is in competition with attractive energy effects: one expects that in this case $h$ needs to
81: be positive for the energy term to overcome the entropic repulsion effect, but quantitative estimates
82: are not a priori obvious.
83: 
84: There is an extensive literature on periodic pinning and wetting models, the majority of which is
85:  restricted
86: to the $T=2$ case, we mention for example \cite{cf:GG,cf:NZ}.
87: \medskip
88:  \item {\it Copolymer near a selective interface.} Much in the same way we introduce
89:  \begin{equation}
90:  \label{eq:copolymer}
91:  \frac{\dd \bP_{N, \go}}{\dd \bP} (S) \, \propto \, \exp \left( \gl \sum _{n=1}^ {N}
92:  \go_n \sign \left( S_n\right) \right),
93:  \end{equation}
94:  where if $S_n=0$ we set $\sign (S_n) := \sign(S_{n-1}) \, \ind_{\{S_{n-1} \neq 0\}}$. This
95:  convention for defining $\sign(0)$, that will be kept throughout the chapter,
96:  has the following simple interpretation: $\sign(S_{n}) = +1,0,-1$ according to whether
97:  the bond joining $S_{n-1}$ and $S_n$ lies above, on,  or below the $x$--axis.
98: 
99:  Also in this case we take a directed walk viewpoint and then
100:  $\bP_{N,\go}$ may be interpreted as a polymeric chain in which the
101:  monomer units, the bonds of the walk, are charged. An interface, the $x$--axis,
102:  separates two solvents, say oil above and water below:
103:  positively charged monomers are hydrophobic and negatively charged ones
104:  are instead hydrophilic.
105: In this case one expects a competition between three possible scenarios: polymer
106: preferring water, preferring oil or undecided between the two and choosing to
107: fluctuate in the proximity of the interface. We will therefore talk of delocalization in water (or oil)
108: or of localization at the interface. Critical cases are of course of particular  interest.
109: 
110: We select \cite{cf:MGO,cf:SD} from the physical literature on periodic copolymers,
111: keeping however in mind that periodic copolymer modeling  has a central
112: role in applied chemistry and material science.
113:  \end{enumerate}
114: 
115: 
116: 
117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
118: 
119: 
120: 
121: \smallskip
122: 
123: \subsection{A general model}
124: We point out that the models presented in \S~\ref{sec:2examples}
125: are particular examples of the polymer measure with Hamiltonian
126: \begin{equation}
127:  \label{eq:genH}
128: \cH _N (S)\, =\,
129:  \sum_{i =\pm 1} \sum _{n=1}^ {N}
130:  \go_n^{(i)} \ind_{\left\{\sign \left( S_n\right)=i\right\}}+
131:  \sum _{n=1}^ {N}
132:  \go^{(0)}_n \ind_{\left\{S_n=0 \right\}}+
133:   \sum _{n=1}^ {N}
134:  \widetilde\go^{(0)}_n \ind_{\left\{\sign\left(S_n\right)=0 \right\}},
135:  \end{equation}
136: where $\go^{(\pm1)}$, $ \go^{(0)}$ and  $\widetilde\go^{(0)}$
137: are periodic sequences of real numbers. Observe that, by our conventions on~$\sign(0)$, the last term gives an energetic contribution (of pinning/depinning type) to the bonds lying on the interface.
138: 
139: Besides being a natural model, generalizing and interpolating between
140: pinning and copolymer models, the general model we consider is the
141: one considered at several instances, see e.g. \cite{cf:SW} and references therein.
142: \smallskip
143: 
144: \begin{rem} \label{rem:1.1}\rm
145: The \textsl{copolymer} case corresponds to $\go^{(+1)}=-\go^{(-1)}= \gl \, \go$ and $ \go^{(0)}=\widetilde\go^{(0)}=0$,
146: while the \textsl{pinning} case corresponds to $\go^{(0)}=\gl \, \go$ and
147: $\go^{(+1)}=\go^{(-1)}=\widetilde\go^{(0)}=0$. We stress
148:  that the \textsl{wetting} case can be included
149: too, with the choice $\go^{(0)}=\gl \go$, $\go^{(-1)}_n=-\infty$ for every $n$ and
150: $\go^{(+1)}=\widetilde\go^{(0)}=0$. Of course plugging $\go^{(-1)}_n=-\infty$ into the Hamiltonian~\eqref{eq:genH}
151:  is a bit formal, but it
152:  simply corresponds to a constraint on $S$ in the polymer measure associated to~$\cH_N$, see \eqref{eq:genP} below. For ease of exposition we will restrict to finite values of the charges~$\go$, but
153:  the generalization is straightforward.
154: \end{rem}
155: \smallskip
156: 
157: 
158: \begin{rem} \rm
159: \label{rem:intper}
160: We take this occasion for stressing that, from an applied viewpoint, the interest  in periodic
161: models of the type we consider appears to be  at least two--fold.
162: On one hand periodic models are often chosen as caricatures of the
163: {\sl quenched disordered} models, like the ones in which the charges are
164: a typical realization of a sequence of independent random variables
165: (e.g. \cite{cf:AS,cf:BdH,cf:G,cf:SW} and references therein).
166: In this respect and taking a mathematical standpoint, the relevance
167: of periodic models, which may be viewed as {\sl weakly inhomogeneous},
168: for understanding the strongly inhomogeneous quenched set--up is
169: at least questionable and the approximation of quenched models
170: with periodic ones, in the limit of large period, poses very interesting and
171: challenging questions. In any case, the precise description
172: of the periodic case that we have obtained in this work
173: highlights limitations and perspectives of periodic modeling
174: for strongly inhomogeneous systems.
175: One the  other hand, as already mentioned above, periodic models
176: are absolutely natural and of direct relevance for application,
177: for example when dealing with {\sl molecularly engineered} polymers~\cite{cf:NN,cf:SD}.
178: \end{rem}
179: 
180: \smallskip
181: 
182: Starting from the Hamiltonian \eqref{eq:genH}, for $a= \rc$ ({\sl constrained}) or
183: $a=\rf$ ({\sl free})
184: we introduce the \textsl{polymer measure}~$\bP^a_{N, \go}$ on $\Z^\N$, defined by
185: \begin{equation}
186: \label{eq:genP}
187: \frac{\dd \bP^a_{N, \go}}{\dd \bP} (S)\, =\,
188: \frac{\exp\left(\cH_N (S)\right)}{\tilde Z_{N, \go}^a} \left(\ind_{\left\{a=\rf\right\} } +
189: \ind_{\left\{a=\rc\right\} } \ind_{\left\{ S_N=0 \right\}}
190: \right),
191: \end{equation}
192: where $\tilde Z_{N, \go}^a := \bE [\exp (\cH_N) \, (\ind_{\{a=\rf\} } +
193: \ind_{\{a=\rc\} } \ind_{\{ S_N=0 \}}) ]$ is the {\sl partition function}, that is the normalization constant.
194: Here $\go$ is a shorthand for the four periodic sequences appearing in the definition~\eqref{eq:genH} of~$\cH_N$, and we will use $T=T(\go)$ to denote the smallest common period of the sequences.
195: 
196: \smallskip
197: 
198: The Laplace asymptotic behavior of $\tilde Z_{N,\go}$
199: plays an important role
200: and the quantity
201: \begin{equation}
202: \label{eq:fe}
203: f_\go \, := \, \lim_{N \to \infty } \frac 1N \log \tilde Z_{N, \go} ^{\rc},
204: \end{equation}
205: is usually called {\sl free energy}. The existence of the limit above follows from
206: a direct super--additivity argument, and it is easy to check that
207: $\tilde Z_{N, \go} ^{\rc}$ can be replaced by $\tilde Z_{N, \go} ^{\rf}$
208: without changing the value of $f_\go$,
209: see e.g.~\cite{cf:G}. The standard free energy approach to this type of models
210: starts from the observation that
211: \begin{equation}
212: \label{eq:fdeloc0}
213: \begin{split}
214: f_\go \, & \;\ge\;
215: \lim_{N\to \infty} \; \frac{1}{N} \, \log \bE
216: \Big[\exp\big(\cH_N(S)\big)\, ; \, S_n>0 \text{ for } n=1, \ldots, N\, \Big]\\
217: & \;=\; \frac 1{T(\go)}\sum_{n=1}^{T(\go)} \go_n^{(+1)} \;+\;
218: \lim_{N\to \infty} \;
219: \frac{1}{N} \, \log \bP\big(S_n>0 \text{ for } n=1, \ldots, N\, \big) \,.
220: \end{split}
221: \end{equation}
222: It is a classical result~\cite[Ch. XII.7]{cf:Fel2} that $\bP (S_n>0 \text{ for } n=1, \ldots, N) \sim c N^{-1/2}$, as~$N\to\infty$,
223: for some $c \in (0,\infty)$ (by
224:  $a_N \sim b_N$ we mean $a_N/b_N \to 1$). Hence the  limit of the last term of \eqref{eq:fdeloc0} is zero and one
225: easily concludes that
226: \begin{equation}
227: \label{eq:fdeloc}
228: f_\go \, \ge\, f_{\go}^{\cD} \, :=\, \max _{i=\pm 1} h_\go(i),
229: \qquad h_\go (i):=\frac 1{T(\go)}\sum_{n=1}^{T(\go)} \go_n^{(i)}
230: .
231: \end{equation}
232: Having in mind the steps in \eqref{eq:fdeloc0}, one is led to the following basic
233: \begin{definition} \label{def:main}
234: The polymer chain defined by \eqref{eq:genP} is said to be:
235: \begin{itemize}
236: \item \textsl{localized (at the interface)} if $f_\go>f_{\go}^{\cD}$;
237: \smallskip
238: \item \textsl{delocalized above the interface} if $f_\go= h_\go (+1)$;
239: \smallskip
240: \item \textsl{delocalized below the interface} if $f_\go= h_\go (-1)$.
241: \end{itemize}
242: \end{definition}
243: \noindent
244: Notice that, with this definition, if $h_\go(+1) = h_\go(-1)$ and the polymer is delocalized, it is delocalized both above and below the interface.
245: 
246: \smallskip
247: 
248: \begin{rem}\rm
249: Observe that the polymer measure~$\bP^a_{N,\go}$ is invariant under the joint transformation $S \to -S$, \ $\go^{(+1)} \to \go^{(-1)}$, hence by symmetry we may (and will) assume that
250: \begin{equation}\label{eq:ass_h}
251: h_\go \, := \, h_\go(+1) \, - \, h_\go(-1) \, \geq \, 0\,.
252: \end{equation}
253: It is also clear that we can add to the Hamiltonian~$\cH_N$ a constant term (with respect to~$S$) without changing the polymer measure. Then we set
254: \begin{align*}
255:     \cH'_N (S) \, &:=\, \cH _N (S) \;-\; \sum _{n=1}^ {N}\go_n^{(+1)}\,,
256: %    &\, = \, \sum _{n=1}^ {N} \big(\go_n^{(-1)}-\go_n^{(+1)}\big) \, \ind_{\left\{\sign \left( S_n\right)=i\right\}}
257: %\;+\; \sum _{n=1}^ {N}  \go^{(0)}_n \, \ind_{\left\{S_n=0 \right\}} \\
258: %    & \phantom{xxxxxxxxxxxxxxxx} + \; \sum _{n=1}^ {N}
259: % \big( \widetilde\go^{(0)}_n -\go_n^{(+1)}\big) \, \ind_{\left\{\sign\left(S_n\right)=0 \right\}}.
260: \end{align*}
261: which amounts to redefining $\go_n^{(+1)} \to 0$, $\go_n^{(-1)} \to (\go_n^{(-1)}-\go_n^{(+1)})$ and $\widetilde\go_n^{(0)} \to (\widetilde\go_n^{(0)}-\go_n^{(+1)})$, and we can write
262: \begin{equation}
263: \label{eq:newP}
264: \frac{\dd \bP^a_{N, \go}}{\dd \bP} (S)\, =\,
265: \frac{\exp\left(\cH_N' (S)\right)}{Z_{N, \go}^a} \left(\ind_{\left\{a=\rf\right\} } +
266: \ind_{\left\{a=\rc\right\} } \ind_{\left\{ S_N=0 \right\}}
267: \right),
268: \end{equation}
269: where $Z_{N, \go}^a$ is a new partition function which coincides with
270: $\tilde Z_{N, \go}^a \exp(- \sum _{n=1}^ {N}\go_n^{(+1)})$. The corresponding free energy~$\tf_\go$ is given by
271: \begin{equation}
272: \label{eq:F}
273:     \tf_\go \;:=\; \lim_{N\to\infty} \frac 1N \log Z_{N,\go}^a \;=\; f_\go - f_\go^\cD\,,
274: \end{equation}
275: and notice that in terms of~$\tf_\go$ the condition for localization (resp. delocalization) becomes $\tf_\go > 0$ (resp.~$\tf_\go = 0$). From now on, speaking of partition function and free energy we will always mean~$Z_{N,\go}^a$ and~$\tf_\go$.
276: \end{rem}
277: 
278: 
279: 
280: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
281: 
282: 
283: 
284: \smallskip
285: 
286: \subsection{From free energy to path behavior}
287: \label{sec:spirit}
288: 
289: In order to understand the spirit of our approach, let us briefly outline our results (complete results are given in~\S~\ref{sec:results} below).
290: 
291: Our first goal is to give necessary and sufficient \textsl{explicit conditions} in terms of the charges~$\go$ for the (de)localization of the polymer chain, see Theorem~\ref{th:as_Z}. We point out that the content of this theorem is in fact much richer, as it gives the \textsl{sharp asymptotic behavior} (and not only the Laplace one \cite{cf:BG}) as~$N\to\infty$ of the constrained partition function~$Z^\rc_{N,\go}$. In particular we show that when the polymer is delocalized ($\tf_\go = 0$) the constrained partition function $Z^\rc_{N,\go}$ is actually vanishing as~$N\to\infty$. Moreover the rate of the decay induces a further distinction in the delocalized regime between a \textsl{strictly delocalized regime} ($Z_{N,\go}^\rc \sim c_1 N^{-3/2}$, $c_1 \in (0, \infty)$) and a \textsl{critical regime} ($Z_{N,\go}^\rc \sim c_2 N^{-1/2}$, $c_2 \in (0, \infty)$).
292: 
293: These asymptotic results are important because they allow to address further interesting issues. For example, it has to be admitted that defining (de)localization in terms of the free energy is not completely satisfactory, because one would like to characterize the
294:  polymer path properties. In different terms, given a polymer measure which is (de)localized according to Definition~\ref{def:main}, to what extent are its typical paths really (de)localized? Some partial answers to this question are known, at least in some particular instances: we mention
295:  here the case of $T(\go)=2$ copolymers \cite{cf:MGO}
296:  and the case of homogeneous pinning and wetting models \cite{cf:DGZ,cf:IY,cf:Upton}.
297: 
298: 
299: Our main aim is to show that, for the whole class of models we are considering, free energy (de)localization does correspond to a \textsl{strong form} of path (de)localization. More precisely, we look at path behavior from two different viewpoints.
300: \smallskip
301: \begin{itemize}
302: \item \emph{Thermodynamic limit.} We show that the measure $\bP_{N,\go}^a$ converges weakly as~$N\to\infty$ toward a measure~$\bP_\go$ on~$\Z^\N$, of which we give an explicit construction, see Section~\ref{sec:infvol}. It turns out that the properties of~$\bP_\go$ are radically different in the three regimes (localized, strictly delocalized and critical), see Theorem~\ref{th:infvol}.
303: It is natural to look at these results as those characterizing the {\sl local} structure of the polymer chain.
304: \item \rule{0pt}{15pt}\emph{Brownian scaling limits.} We prove that the diffusive rescaling of the polymer measure $\bP_{N,\go}^a$ converges weakly in $C([0,1])$ as~$N\to\infty$. Again the properties of the limit process, explicitly described in Theorem~\ref{th:scaling}, differ considerably in the three regimes.
305: Moreover we stress that scaling limits describe {\sl global} properties of the chain.
306: \end{itemize}
307: \smallskip
308: We insist on the fact  that the path analysis just outlined has been obtained exploiting heavily the sharp asymptotic behavior of~$Z_{N,\go}^\rc$ as~$N\to\infty$.
309: In this sense our results are the direct sharpening of the Large Deviations
310: approach taken in \cite{cf:BG}, where a formula  for $\tf_\go$
311: was obtained for periodic copolymers (but the method of course directly extends to the general
312: case considered here).
313: Such a formula (see \S~\ref{sec:as_Z_loc}), that reduces the problem of computing the free energy
314: to a finite dimensional problem connected to a suitable
315: Perron--Frobenius matrix, in itself suggests the new approach taken here
316: since it makes rather apparent the link between periodic
317: copolymers and the class of {\sl Markov renewal processes}~\cite{cf:Asm}.
318: On the other hand, with respect to \cite{cf:BG}, we leave aside any issue
319: concerning the phase diagram (except for \S~\ref{sec:P} below).
320: 
321: 
322: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
323: 
324: 
325: \smallskip
326: 
327: \subsection{The order parameter~$\gd^\go$}
328: \label{sec:order_par}
329: 
330: It is a remarkable fact that the dependence of our results on the charges~$\go$ is essentially encoded in one single parameter~$\gd^\go$, that can be regarded as the \textsl{order parameter} of our models. For the definition of this parameter, we need some preliminary notation. We start with the law of the first return to zero of the original walk:
331: \begin{equation} \label{eq:first_ret}
332:     \tau_1 := \inf \{n>0:\ S_n = 0\} \qquad \quad K(n) \;:=\; \bP \big( \tau_1 = n \big)\,.
333: \end{equation}
334: It is a classical result \cite[Ch. XII.7]{cf:Fel2} that
335: \begin{equation} \label{eq:as_K}
336:     \exists \lim_{n\to\infty} n^{3/2} \, K(n) \;=:\; c_K \in (0,\infty)\,.
337: \end{equation}
338: Then we introduce the Abelian group $\bbS:= \Z/ (T\Z)$ and to indicate that an integer~$n$ is in the equivalence class $\gb\in\bbS$ we write equivalently $[n]=\gb$ or $n\in\gb$. Notice that the charges~$\go_n$ are functions of~$[n]$, and with some abuse of notation we can write~$\go_{[n]}:=\go_n$. The key observation is that, by the $T$--periodicity of the charges~$\go$ and by the definition~\eqref{eq:ass_h} of~$h_\go$, we can write
339: \begin{equation*}
340:     \sum_{n=n_1+1}^{n_2} (\go_n ^{(-1)}-\go_n ^{(+1)}) \;=\; -(n_2-n_1) \, h_\go \;+\; \Sigma_{[n_1],[n_2]}\,.
341: \end{equation*}
342: Thus we have decomposed the above sum into a drift term and a more fluctuating term, where the latter has the remarkable property of depending on~$n_1$ and~$n_2$ only through their equivalence classes~$[n_1]$ and~$[n_2]$. Now we can define three basic objects:
343: \begin{itemize}
344: \smallskip
345: \item for $\ga, \gb\in \bbS$ and $\ell \in \N $ we set
346: \begin{equation} \label{eq:def_Phi}
347: \Phi^\go_{\ga, \gb}(\ell)\, :=\,
348: \begin{cases}
349:     \go ^{(0)}_{\gb} \;+\; \Big(\tilde  \go ^{(0)}_{\gb} - \go^{(+1)}_\gb \Big) & \text{if }
350:     \ell=1,\ \ell \in \gb-\ga \\
351:      \go ^{(0)}_{\gb} +
352:     \rule{0pt}{22pt} \displaystyle  \log  \bigg[ \frac 12 \Big(1+ \exp\big( -\ell \, h_\go  + \Sigma_{\ga,\gb}\big)\Big) \bigg] &\text{if } \ell > 1,\ \ell \in \gb-\ga\\
353:     \rule{0pt}{16pt} 0 & \text{otherwise}
354: \end{cases}
355: \;,
356: \end{equation}
357: which is a sort of integrated version of our Hamiltonian;
358: 
359: \medskip
360: \item for $x\in\N$ we introduce the $\bbS \times \bbS$ matrix $M^\go_{\ga,\gb}(x)$ defined by
361: \begin{equation} \label{eq:matrix_cgz}
362: M^\go_{\ga,\gb}(x) \;:=\; e^{\Phi^\go_{\ga,\gb}(x)} \, K(x) \, \ind_{(x \in \gb -\ga)}\,;
363: \end{equation}
364: 
365: \medskip
366: \item summing the entries of~$M^\go$ over~$x$ we get a~$\bbS\times\bbS$ matrix that we call~$B^\go$:
367: \begin{equation} \label{eq:def_B}
368:     B^\go_{\ga,\gb} := \sum_{x\in\N} M^\go_{\ga,\gb}(x)\,.
369: \end{equation}
370: \end{itemize}
371: \smallskip
372: The meaning and motivation of these definitions, that at this point might appear artificial,
373: are explained in detail in \S~\ref{sec:RWexcurs}. For the moment we only stress that the
374: above quantities are \textsl{explicit functions} of the charges~$\go$ and of the law of the
375: underlying random walk (to lighten the notation, the $\go$--dependence of these quantities will
376: be often dropped in the following).
377: 
378: \smallskip
379: We can now define our order parameter~$\gd^\go$. Observe that $B_{\ga,\gb}$ is a finite dimensional matrix with nonnegative entries, hence the Perron--Frobenius (P--F) Theorem (see e.g. \cite{cf:Asm}) entails that $B_{\ga,\gb}$ has a unique real positive eigenvalue, called the Perron--Frobenius eigenvalue, with the property that it is a simple root of the characteristic polynomial and that it coincides with the spectral radius of the matrix. This is exactly our parameter:
380: \begin{equation} \label{eq:def_delta}
381: \gd^\go \ := \ \text{Perron--Frobenius eigenvalue of $B^\go$}\,.
382: \end{equation}
383: 
384: \smallskip
385: 
386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
387: 
388: \subsection{The main results}
389: \label{sec:results}
390: 
391: Now we are ready to state our results. We start characterizing the (de)localization of the polymer chain in terms of~$\gd^\go$.
392: 
393: %For notational convenience, here we give the asymptotic behavior of~$Z_{N,\go}^\rc$ only as~$N\to\infty$ with $[N]=0$, that is along systems sizes that are multiples of the charges period~$T(\go)$ (the reader is referred to Section~\ref{sec:as_Z} for the general case).
394: 
395: \medskip
396: 
397: \begin{theorem}[\bf Sharp asymptotics] \label{th:as_Z}
398: The polymer chain is localized if and only if $\gd^\go > 1$.
399: More precisely, the asymptotic behavior of $Z_{N,\go}^\rc$ as
400: $N\to\infty$, $[N]=\eta$ is given by
401: \smallskip
402: \begin{enumerate}
403: \item for $\gd^\go > 1$ (\textsl{localized regime}) \ $Z_{N,\go}^\rc \;\sim\; C^>_{\go,\eta} \, \exp \big(\tf_\go N \big)$ \,;
404: \medskip
405: \item for $\gd^\go < 1$ (\textsl{strictly delocalized regime}) \ $Z_{N,\go}^\rc \;\sim\; C^<_{\go,\eta} \, / \, N^{3/2}$ \,;
406: \medskip
407: \item for $\gd^\go = 1$ (\textsl{critical regime}) \ $Z_{N,\go}^\rc \;\sim\; C^=_{\go,\eta} \, / \, \sqrt{N} $ \,,
408: \end{enumerate}
409: \smallskip
410: where $\tf_\go > 0$ is the free energy and its explicit definition
411: in terms of~$\go$ is given in~\S~\ref{sec:as_Z_loc}, while~$C^>_{\go,\eta}$,
412: $C^<_{\go,\eta}$ and $C^=_{\go,\eta}$ are
413: explicit positive constants, depending on $\go$ and $\eta$,
414: whose value is given in Section~\ref{sec:as_Z}.
415: \end{theorem}
416: 
417: \medskip
418: \begin{rem}
419: \label{sec:compdis}
420: \rm
421: Theorem~\ref{th:as_Z}
422: is the building block of all the path analysis that follows. It is therefore
423: important to stress that, in the quenched disordered case,
424: cf. Remark~\ref{rem:intper},  such a strong statement in general
425: does not hold, see \cite[Section 4]{cf:GT}.
426: \end{rem}
427: \medskip
428: 
429: Next we investigate the thermodynamic limit, that is the
430: weak limit as~$N\to\infty$ of the sequence of
431: measures~$\bP^{a}_{N,\go}$ on~$\Z^\N$ (endowed with the
432: standard product topology). The next theorem provides a
433: first connection between free energy (de)localization
434: and the corresponding path properties.
435: 
436: Before stating the result, we need a notation: we denote
437: by $\cP$ the set of $\omega$ such that:
438: %
439: \begin{equation}\label{sit1}
440: \cP \, := \, \left\{ \go \, : \ \gd^\go \leq 1, \quad
441: h_\go=0, \quad \exists \ \ga,\gb: \ \Sigma_{\ga,\gb} \ne 0
442: \right\},
443: \end{equation}
444: %
445: \[
446: \cP^< \, := \, \cP \cap\{\gd^\go < 1\},
447: \qquad \cP^= \, := \, \cP \cap\{\gd^\go = 1\}.
448: \]
449: Here $\cP$ stands for {\sl problematic}, or {\sl pathologic}.
450: Indeed, we shall see that for $\go\in\cP$ the results are
451: weaker and more involved than for $\go\notin\cP$. We stress however
452: that these restrictions do not concern localized regime,
453: because $\cP\subset\{\go:\delta^\go\leq 1\}$. We also notice that
454: for the two motivating models of \S~\ref{sec:2examples}, the pinning
455: and the copolymer models, $\go$ {\it never} belongs to
456: $\cP$. This is clear for the pinning case,
457: where by definition $\Sigma\equiv 0$. On the other hand, in the copolymer case
458: it is known that if $h_\go=0 $ and $\exists \ \ga,\gb: \ \Sigma_{\ga,\gb} \ne 0$
459: then $\delta^\go>1$: see \S~\ref{app:loc} or~\cite{cf:BG}.
460: In reality
461:  the {\sl pathological} aspects
462: observed for $\go \in \cP$
463: may be understood in statistical mechanics terms
464: and we sketch an interpretation  in   \S~\ref{sec:P} below:
465: this goes rather far from the point of view adopted here,
466: since it is an issue
467: tightly entangled with the analysis of the free energy.
468: It will therefore be taken up in a further work.
469: 
470: 
471: 
472: \medskip
473: 
474: \begin{theorem}[\bf Thermodynamic limit]\label{th:infvol}
475: If $\go\notin\cP^<$, then
476: both the polymer measures $\bP^\rf_{N, \go}$ and
477: $\bP^\rc_{N, \go}$ converge as $N\to\infty$ to the same limit
478: $\bP_\go$, law of an irreducible Markov process on $\Z$ which is:
479: \smallskip
480: \begin{enumerate}
481: %
482: \item positive recurrent if $\gd^\go>1$ (\textsl{localized regime})\,;
483: %
484: \medskip
485: \item transient if $\gd^\go<1$
486: (\textsl{strictly delocalized regime})\,;
487: %
488: \medskip
489: \item null recurrent if $\gd^\go=1$
490: (\textsl{critical regime})\,.
491: %
492: \end{enumerate}
493: \smallskip
494: %
495: If $\go\in\cP^<$ (in particular $\gd^\go < 1$), for all $\eta\in\bbS$ and $a=\rf,\rc$ the measure
496: $\bP^a_{N, \go}$ converges as $N\to\infty$, $[N]=\eta$ to
497: $\bP^{a,\eta}_{\go}$, law of an irreducible transient Markov
498: chain on $\Z$.
499: \end{theorem}
500: 
501: \medskip\noindent
502: We stress that in all regimes the limit law~$\bP_\go$ or
503: $\bP^{a,\eta}_{\go}$ has an explicit
504: construction in terms of~$M^\go_{\ga,\gb}(x)$, see
505: Section~\ref{sec:infvol} for details.
506: 
507: %\medskip\noindent
508: %In fact, $\go\in\cP$ means that:
509: %
510: %\begin{itemize}
511: %
512: %\item on one hand, $h_\go=0$ but $\Sigma$ is not identically zero,
513: %so that the ``copolymer'' part of the Hamiltonian alone would
514: %lead to localization
515: %
516: %\item on the other hand, we are in the delocalized regime
517: %$\delta^\go<1$, and this is necessarily due to
518: %the ``non-copolymer'' part of $\cH_N$.
519: %
520: %\end{itemize}
521: %
522: 
523: \bigskip
524: 
525: We finally turn to the analysis of the \textsl{diffusive rescaling}
526: of the polymer measure~$\bP_{N,\go}^a$. More precisely, let us
527: define the map $X^N: {\mathbb R}^N \mapsto C([0,1])$:
528: \[
529: X^N_t(x) \, = \, \frac{ x_{\lfloor Nt\rfloor}}{\sigma N^{1/2}} + (Nt-\lfloor Nt\rfloor) \,
530: \frac {x_{\lfloor Nt\rfloor+1}-x_{\lfloor Nt\rfloor}}{\sigma N^{1/2}}, \qquad t\in[0,1],
531: \]
532: where $\lfloor \,\cdot\, \rfloor$ denotes the integer part and $\gs^2 := 2p$ is the variance
533: of~$X_1$ under the original random walk measure~$\bP$.
534: Notice that $X^N_t(x)$ is nothing but the linear interpolation of~$\{x_{\lfloor Nt \rfloor}/(\gs\sqrt{N})\}_{t \in \frac{\N}{N} \cap [0,1]}$. For $a=\rf,\rc$ we set:
535: \[
536: Q^{a}_{N,\go} \, := \, {\bf P}^{a}_{N,\go} \circ (X^N)^{-1},
537: \]
538: %consider the linear interpolation of the process $\{S_{tN}/(\gs \sqrt{N})\}$
539: %for $t \in [0,1] \cap \Z/N$ under the measure~$\bP_{N,\go}^a$
540: Then $Q_{N,\go}^a$ is a measure on $C([0,1])$, the space of real
541: continuous functions defined on the interval~$[0,1]$, and we want
542: to study the behavior as~$N\to\infty$ of this sequence of measures.
543: 
544: \smallskip
545: 
546: We start fixing a notation for the following standard processes:
547: \smallskip
548: \begin{itemize}
549: \item the Brownian motion $\left\{B_\tau\right\}_{\tau\in [0,1]}$;
550: \smallskip
551: \item the Brownian bridge $\left\{\beta_\tau\right\}_{\tau\in
552: [0,1]}$ between $0$ and $0$;
553: \smallskip
554: \item the Brownian motion {\sl conditioned to stay
555: non-negative on $[0,1]$} or, more precisely, the Brownian meander
556: $\{m_{\tau}\}_{\tau\in[0,1]}$, see \cite{cf:RevYor};
557: %
558: \smallskip
559: \item the Brownian bridge {\sl conditioned to stay
560: non-negative on $[0,1]$} or, more precisely, the normalized
561: Brownian excursion $\{e_{\tau}\}_{\tau\in[0,1]}$, also known as the
562: Bessel bridge of dimension 3 between $0$ and $0$, see
563: \cite{cf:RevYor} .
564: \end{itemize}
565: \smallskip
566: Then we introduce a modification of the above processes labeled by a
567: parameter $p \in [0,1]$:
568: %
569: \begin{itemize}
570: \item \rule{0pt}{12pt}the process $\{B^{(p)}_\tau\}_{\tau \in [0,1]}$
571: is the so--called {\sl skew Brownian motion of parameter~$p$},
572: cf.~\cite{cf:RevYor}. More explicitly, $B^{(p)}$ is a process such
573: that~$|B^{(p)}| = |B|$ in distribution, but in which the sign of each
574: excursion is chosen to be $+1$ (resp.~$-1$) with probability~$p$
575: (resp.~$1-p$) instead of~$1/2$. In the same way, the process
576: $\{\gb^{(p)}_\tau\}_{\tau \in [0,1]}$ is the skew Brownian bridge
577: of parameter~$p$. Notice that for $p=1$ we have $B^{(1)}=|B|$ and
578: $\beta^{(1)}=|\beta|$ in distribution.
579: \smallskip
580: \item the process $\{m^{(p)}_\tau\}_{\tau \in [0,1]}$ is defined by
581: \[
582: \bbP(m^{(p)}\in dw) \, := \, p \, \bbP(m\in dw) \, + \,
583: (1-p) \, \bbP(-m\in dw),
584: \]
585: i.e. $m^{(p)}=\sigma  m$, where $\bbP(\sigma=1)=1-\bbP(\sigma=-1)=p\,$
586: and $(m,\sigma)$ are independent. The process
587: $\{e^{(p)}_\tau\}_{\tau \in [0,1]}$ is defined in exactly the same manner.
588: For $p=1$ we have $m^{(1)}=m$ and $e^{(1)}=e$.
589: \end{itemize}
590: %
591: \smallskip
592: Finally, we introduce a last process, labeled by two
593: parameters $p,q\in[0,1]$:
594: %
595: \begin{itemize}
596: \item \rule{0pt}{12pt}consider a r.v. $U\mapsto [0,1]$ with
597: the arcsin law: $\bbP(U\leq t)=\frac 2\pi\, \arcsin\sqrt t$,
598: and processes $\gb^{(p)}$, $m^{(q)}$ as defined above,
599: with $(U,\gb^{(p)},m^{(q)})$ independent triple.
600: Then we denote by $\{B^{(p,q)}_\tau\}_{\tau \in [0,1]}$
601: the process defined by:
602: \[
603: B^{(p,q)}_\tau \, := \,
604: \begin{cases}
605: \sqrt U \, \gb^{(p)}_{\frac\tau U}
606: & \text{if } \tau \leq U
607: \\
608: \rule{0pt}{24pt}
609: \sqrt{1-U} \, m^{(q)}_{\frac{\tau-U}{1-U}}
610: & \text{if } \tau > U
611: \end{cases} \;.
612: \]
613: Notice that the process $B^{(p,q)}$ differs from the $p$--skew Brownian motion
614: $B^{(p)}$ only for the last excursion in~$[0,1]$, whose sign is~$+1$ with
615: probability~$q$ instead of~$p$.
616: \end{itemize}
617: %
618: 
619: 
620: 
621: 
622: \medskip
623: 
624: We are going to show that the sequence $\{Q_{N,\go}^a\}$ has a
625: weak limit as~$N\to\infty$ (with a weaker statement if $\go\in\cP$).
626: Again the properties of the limit process differ considerably in
627: the three regimes $\gd^\go>1$, $\gd^\go<1$ and $\gd^\go=1$.
628: However for the precise description of the limit processes, for the regimes
629: $\gd^\go = 1$ and~$\gd^\go < 1$ we need to distinguish between
630: $a\in\{\rf,\rc\}$ and to introduce
631: further parameters~$\tt p_\go, \tt q_\go$, defined as follows:
632: \begin{itemize}
633: \item case $\gd^\go=1$\, :
634: \begin{itemize}
635: \item $\tt p_\go := \tt p_\go^=$,  defined in \eqref{eq:p_om3}. We point out two special cases:
636: if~$h_\go>0$ then~$\tt p^=_\go =1$, while if~$h_\go=0$
637: and~$\Sigma \equiv 0$ then~$\tt p^=_\go = 1/2$;
638: %
639: \item for each $\eta\in\bbS$,
640: $\tt q_\go := \tt q_{\go,\eta}^=$, defined by \eqref{eq:p_om4}.
641: \end{itemize}
642: %
643: %\item $\tt p_\go := \tt p_\go^=$,  defined in \eqref{eq:p_om3}. We point out two special cases:
644: %if~$h_\go>0$ then~$\tt p^=_\go =1$, while if~$h_\go=0$
645: %and~$\Sigma \equiv 0$ then~$\tt p^=_\go = 1/2$;
646: %
647: %\item for each $\eta\in\bbS$,
648: %$\tt q_\go := \tt q_{\go,\eta}^=$, defined by \eqref{eq:p_om4}.
649: %\end{itemize}
650: \item case $\gd^\go<1$\, :
651: \begin{itemize}
652: \item
653:  $\go \notin \cP^<$:\: if~$h_\go>0$ we set
654: $\tt p_\go:=\tt p_\go^<:=1$
655: while if~$h_\go = 0\,$ we set $\tt p_\go:=\tt p_\go^< := 1/2$;
656: \item  $\go \in \cP^<$:\: for each $\eta\in\bbS$
657: and $a=\rf,\rc$,
658: ${\tt p}_\go:={\tt p}_{\go, \eta}^{<, a}$ is defined in \eqref{eq:p_om}
659: and \eqref{eq:p_om2}.
660: \end{itemize}
661: \end{itemize}
662: 
663: \medskip
664: %
665: \begin{theorem}[\bf Scaling limits]
666: \label{th:scaling}
667: %
668: If $\go\notin\cP$, then
669: the sequence of measures~$\{Q_{N,\go}^a\}$ on $C([0,1])$ converges weakly
670: as~$N\to\infty$. More precisely:
671: \smallskip
672: \begin{enumerate}
673: %
674: \item for $\delta^\go >1$ (\textsl{localized regime}) $Q^a_{N,\go}$ converges to the measure
675: concentrated on the constant function taking the value zero\,;
676: %
677: \medskip
678: \item for $\delta^\go<1$ (\textsl{strictly delocalized regime}):
679: %
680: \begin{itemize}
681: \item
682: $Q^\rf_{N,\go}$ converges to the law of $m^{(\rm p_\go^<)}$\,;
683: %
684: \item
685: $Q^\rc_{N,\go}$ converges to the law of $e^{(\rm p_\go^<)}$\,;
686: %
687: \end{itemize}
688: %
689: \medskip
690: \item for $\delta^\go=1$ (\textsl{critical regime}):
691: %
692: \begin{itemize}
693: \item
694: $Q^\rf_{N,\go}$ converges to the law of $B^{{(\rm p_\go^=)}}$\,;
695: %
696: \item $Q^\rc_{N,\go}$ converges to the law of $\beta^{(\rm p_\go^=)}$\,.
697: \end{itemize}
698: \smallskip
699: \end{enumerate}
700: %
701: If $\go\in\cP$, then for all $\eta\in\bbS$ the measures
702: $Q^\rc_{N, \go}$ and $Q^\rf_{N, \go}$ converge as $N\to\infty$,
703: $[N]=\eta$ to, respectively:
704: \begin{enumerate}
705: %
706: \item for $\delta^\go<1$,
707: the law of $e^{{(\tt p_{\go,\eta}^{<,\rc})}}$ and $m^{{(\tt p_{\go,\eta}^{<,\rf})}}$.
708: %
709: \item for $\delta^\go=1$,
710: the law of $\beta^{{(\tt p_{\go}^{=})}}$ and $B^{{(\tt p_\go^=,\tt q_{\go,\eta}^=)}}$.
711: %
712: \end{enumerate}
713: %
714: \end{theorem}
715: %
716: 
717: 
718: 
719: 
720: \medskip
721: 
722: Results on thermodynamic limits in the direction of Theorem \ref{th:infvol}
723: have been obtained in the physical literature by exact computations either for homogeneous
724: polymers or for $T=2$ pinning models and copolymers, see e.g. \cite{cf:MGO},
725: while Brownian scaling limits have been heuristically derived
726: at several instances, see e.g. \cite{cf:Upton}.
727: Rigorous results corresponding to our three main theorems have been obtained
728: for homogeneous pinning/wetting models
729: in \cite{cf:DGZ,cf:IY}. We would like to stress the very much richer variety
730: of limit processes that we have obtained in our general context.
731: 
732: \smallskip
733: 
734: \subsection{About the regime $\cP$}
735: \label{sec:P}
736: We have seen, cf. Theorem~\ref{th:infvol},
737: that if $\go \in \cP^<$ the infinite volume limit (in particular
738:  the probability that the walk
739: escapes either to $+\infty$ or to $-\infty$) depends on
740: $a=\rc$ or $\rf$ and on the subsequence $[N]=\eta \in \bbS$.
741:  This reflects  directly into
742: Theorem~\ref{th:scaling} and  in this case also the $\cP^=$ regime is affected,
743: but only for $a=\rf$ and the change is restricted to the
744: sign of the very last excursion of the process.
745: It is helpful to keep in mind that $\go \in \cP$ if and only if
746: there is a non trivial unbiased copolymer part, that is $h_\go=0$ but
747: the matrix $\gS$ is non trivial, and at the same time the polymer is delocalized.
748: It is known (\S~\ref{app:loc} and \cite{cf:BG}) that in absence of
749:  pinning terms, that is $\go^{(0)}_n= \tilde \go^{(0)}_n=0$ for every $n$,
750:   the polymer is localized. However if the pinning rewards are
751:  sufficiently large and negative, one easily sees that (de)pinning takes over
752:  and the polymer delocalizes. This is the phenomenon that characterizes
753:  the regime $\cP$ and its lack of uniqueness of limit measures.
754: \smallskip
755: 
756: 
757: 
758: Lack of uniqueness of infinite volume measures and   dependence
759: on  boundary conditions do not come as a surprise if one takes
760: a statistical mechanics viewpoint and if one notices that the system undergoes
761: a {\sl first order} phase transition exactly at $\cP $. In order to be more precise
762: let us consider the particular case of
763: \begin{equation}
764: \label{eq:Pfig}
765:  \frac{\dd \bP_{N, \go}}{\dd \bP} (S) \, \propto \, \exp \left( \sum _{n=1}^ {N}\left(
766:  \go_n +h \right)\sign \left( S_n\right) -\gb \sum _{n=1}^ {N}
767:   \ind_{\{ S_n=0\}}\right),
768: \end{equation}
769: with $h$ and $\gb$ two real parameters and $\go$ a fixed non trivial centered
770: ($\sum_{n=1}^T\go_n=0$) periodic configuration of charges.
771: The phase diagram of such a model is sketched in Figure~\ref{fig:P}.
772: In particular it is easy to show that for $h=0$ and  for $\gb$ large and positive
773: the polymer is delocalized and, recalling that for $\gb=0$
774: the polymer is localized, by monotonicity of the free energy in $\gb$ one immediately
775: infers that there exists $\gb_c>0$ such that localization prevails for $\gb<\gb_c$,
776: while the polymer is delocalized (both above and below the interface)
777: if $\gb\ge \gb_c$. However the two regimes of delocalization above
778: or below the interface, appearing for example
779: as soon as $h$ is either positive or negative and $\gb \ge \gb_c$,
780: are characterized by opposite values ($\pm 1$) of $\varrho=\varrho (h,\gb):=
781: \lim_{N\to \infty}\bE_{N, \go}
782: \left[ N^{-1}\sum _{n=1}^ {N}
783:  \sign \left( S_n\right) \right]$ and of course $\varrho$ is the derivative of the
784:  free energy with respect to $h$. Therefore the free energy is not
785:  differentiable at $h=0$ and we say that there is a {\sl first order phase transition}.
786: First order phase transitions are usually associated to multiple infinite
787: volume limits ({\sl phase coexistence}).
788: A detailed analysis of this interesting phenomenon
789: will be given elsewhere.
790: 
791: \begin{figure}[h]
792: \begin{center}
793: \leavevmode
794: \epsfysize =7 cm
795: %\epsfxsize =10 cm
796: \psfragscanon
797: \psfrag{D}[c][c]{\Large $\cD$}
798: \psfrag{L}[c][c]{ \Large $\cL$}
799: \psfrag{r1}[c][c]{$\varrho=+1$}
800: \psfrag{r2}[c][c]{ $\varrho=-1$}
801: \psfrag{0}[c][c]{$0$}
802: \psfrag{h}[c][c]{ $h$}
803: \psfrag{bc}[c][c]{$\gb _c$}
804: \psfrag{b}[c][c]{$\gb$}
805: \epsfbox{p.eps}
806: \end{center}
807: \caption{\label{fig:P}
808: A sketch of the phase diagram for  the model \eqref{eq:Pfig}.
809: In this case, with abuse of notation, $\cP=\{ (h, \gb): \, h=0, \, \gb \ge \gb_c\}$.
810: Approaching $\cP$ in the sense of the dashed arrowed lines one
811: observes the two sharply different behaviors of paths completely delocalized
812: above ($\varrho=+1$) or below ($\varrho=-1$) the interface.
813: }
814: \end{figure}
815: 
816: 
817: 
818: 
819: 
820: \smallskip
821: 
822: \subsection{Outline of the exposition}
823: 
824: In Section~\ref{sec:as_Z} we study the asymptotic behavior of $Z^\rc_{N,\go}$,
825: proving Theorem~\ref{th:as_Z}. In Section 3 we compute the thermodynamic
826: limits of $\bP_{N,\go}^a$, proving Theorem~\ref{th:infvol}. In Section
827: 4 we compute the scaling limits of $\bP_{N,\go}^a$, proving Theorem~\ref{th:scaling}.
828: Finally, in Section~\ref{app_cgz} we give the proof of some technical
829: results and some additional material.
830: 
831: 
832: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
833: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
834: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
835: 
836: \smallskip
837: \section{Sharp asymptotic behavior of the partition function}
838: \label{sec:as_Z}
839: 
840: In this section we are going to derive the precise asymptotic behavior of~$Z_{N,\go}^\rc$, in particular proving Theorem~\ref{th:as_Z}. The key observation is that the study of the partition function for the models we are considering can be set into the framework of the theory of \textsl{Markov renewal processes}, see~\cite[Ch.~VII.4]{cf:Asm}. We start recalling the basic notions of this theory and setting the relative notation.
841: 
842: \smallskip
843: \subsection{Markov Renewal Theory}
844: 
845: Given a finite set $\bbS$ (for us it will always be $\Z/(T\Z)$), by a \textsl{kernel} we mean a family of nonnegative $\bbS\times\bbS$ matrices $F_{\ga,\gb}(x)$ depending on a parameter $x\in\N$. We say that the kernel $F_{\ga,\gb}(x)$ is \textsl{semi--Markov} if $F_{\ga,\cdot}(\cdot)$ is a probability mass function on $\bbS\times\N$ for every $\ga\in\bbS$, that is if $\sum_{\gb,x}F_{\ga,\gb}(x)=1$.
846: 
847: \smallskip
848: 
849: A semi--Markov kernel~$F_{\ga,\gb}(x)$ has a simple probabilistic interpretation: it defines a Markov chain $\{(J_k,T_k)\}$ on $\bbS\times\N$ through the transition kernel given by
850: \begin{equation} \label{eq:trans_kernel}
851: \bbP\big[(J_{k+1},T_{k+1}) = (\gb,x) \:\big|\: (J_{k},T_{k}) = (\ga,y) \big] = F_{\ga,\gb}(x)\,.
852: \end{equation}
853: In this case we say that the process $\{J_k,T_k\}$ is a (discrete) \textsl{Markov--renewal process}, the~$\{T_k\}$ being thought of as interarrival times. This provides a generalization of classical renewal processes, since the $\{T_k\}$ are no longer IID but their laws are rather \textsl{modulated} by the process~$\{J_k\}$. Since the r.h.s. of \eqref{eq:trans_kernel} does not depend on~$y$, it follows that $\{J_k\}$ is a Markov chain, and it is called the \textsl{modulating chain} of the Markov renewal process (observe that in general the process $\{T_k\}$ is \textsl{not} a Markov chain). The transition kernel of~$\{J_k\}$ is given by $\sum_{x\in\N} F_{\ga,\gb}(x)$. We will assume that this chain is irreducible (therefore positive recurrent, since $\bbS$ is finite) and we denote by $\{\nu_\ga\}_{\ga\in\bbS}$ its invariant measure.
854: 
855: \smallskip
856: 
857: Given two kernels $F$ and $G$, their convolution $F*G$ is the kernel defined by
858: \begin{equation} \label{eq:convolution}
859: (F*G)_{\ga,\gb}(x) := \sum_{y\in\N} \sum_{\gamma\in\bbS} F_{\ga,\gamma}(y) G_{\gamma,\gb}(x-y) = \sum_{y\in\N} \big[ F(y) \cdot G(x-y) \big]_{\ga,\gb} \,,
860: \end{equation}
861: where $\cdot$ denotes matrix product. Observe that if $F$ and $G$ are semi--Markov kernels, then $F*G$ is semi--Markov too. With standard notation, the $n$--fold convolution of a kernel~$F$ with itself will be denoted by~$F^{*n}$, the $n=0$ case being by definition the identity kernel $[F^{*0}]_{\ga,\gb}(x):=\ind_{(\gb=\ga)} \ind_{(x=0)}$.
862: 
863: \smallskip
864: 
865: A fundamental object associated to a semi--Markov kernel $F$  the so--called \textsl{Markov--Green function} (or Markov--renewal kernel), which is the kernel $\bs U$ defined by
866: \begin{equation} \label{eq:renewal_kernel}
867: \bs{U}_{\ga,\gb}(x) := \sum_{k=0}^\infty \big[ F^{*k} \big]_{\ga,\gb}(x)\,.
868: \end{equation}
869: Of course the kernel $\bs U$ is the analog of the Green function of a classical renewal process, and it has a similar probabilistic interpretation in terms of the associated Markov renewal process~$\{(J_k,T_k)\}$:
870: \begin{equation}\label{eq:prob_interpret}
871:     \bs U_{\ga,\gb} (x) = \bbP_\ga \big[ \exists k \ge 0 :\; T_0+\ldots +T_k=x\;,\; J_{k} = \gb \big]\,,
872: \end{equation}
873: where $\bbP_\ga$ is the law of $\{(T_k,J_k)\}$ conditioned on $\{J_0=\ga, T_0 = 0\}$.
874: 
875: \smallskip
876: 
877: We need some notation to treat our periodic setting: we say that a kernel~$F_{\ga,\gb}(x)$ has period $T\in\N$ if the set $\{x: \bs{U}_{\ga,\ga}(x) \neq 0\}$ is contained in $T\Z$, for the least such~$T$ (this definition does not depend on $\ga$ because the chain $\{J_k\}$ is supposed to be irreducible, see the discussion at p.~208 of \cite{cf:Asm}). It follows that the set $\{x: \bs{U}_{\ga,\gb}(x) \neq 0\}$ is contained in the translated lattice $\gamma(\ga,\gb) + T\N$, where $\gamma(\ga,\gb)\in\{0,\ldots,T-1\}$ (for us it will be $\gamma(\ga,\gb) = [\gb-\ga]$).
878: 
879: \smallskip
880: 
881: In analogy to the classical case, the asymptotic behavior of $\bs U_{\ga,\gb}(x)$
882: as $x\to\infty$ is of particular interest. Let us define the (possibly infinite)
883: \textsl{mean} $\mu$ of a semi--Markov kernel $F_{\ga,\gb}(x)$ as
884: \begin{equation} \label{eq:mu}
885: \mu := \sum_{\ga,\gb \in \bbS} \sum_{x\in\N} x\, \nu_\ga\, F_{\ga,\gb}(x) \,.
886: \end{equation}
887: Then we have an analog of Blackwell's Renewal Theorem, that in our periodic setting reads as
888: \begin{equation} \label{eq:per_renewal_theorem}
889: \exists \limtwo{x\to\infty}{[x] = \gb-\ga} \bs{U}_{\ga,\gb}(x) \;=\; T \frac{\nu_\gb}{\mu}\,,
890: \end{equation}
891: cf. Corollary~2.3 p.~10 of \cite{cf:Asm} for the classical case.
892: 
893: \smallskip
894: 
895: We will see that determining the asymptotic behavior of $\bs U_{\ga,\gb}(x)$ when the kernel $F_{\ga,\gb}(x)$ is no more semi--Markov is the key to get the asymptotic behavior of the partition function~$Z_{N,\go}^\rc$.
896: 
897: 
898: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
899: 
900: 
901: \smallskip
902: \subsection{A random walk excursion viewpoint}
903: \label{sec:RWexcurs}
904: 
905: Now we are ready to make explicit the link between the partition function for our model and the Theory of Markov Renewal Processes. Let us look back to our Hamiltonian \eqref{eq:genH}: its specificity comes from
906: the fact that it can be decomposed in an efficient way by considering the
907: return times to the origin of $S$. More precisely we set for~$j \in \N$
908: \begin{equation*}
909:     \tau_0=0 \qquad \quad \tau_{j+1}=\inf\{ n>\tau_j:\, S_n=0 \}\,,
910: \end{equation*}
911: and for $\bP$--typical trajectories of $S$ one has an infinite sequence $\tau:=\{\tau_j\}_j$ of stopping
912: times. We set $T_j= \tau_j-\tau_{j-1}$ and of course
913: $\{ T_j\}_{j=1, 2, \ldots}$ is, under~$\bP$,  an IID sequence.
914: By conditioning on $\tau$ and integrating on the up--down symmetry
915: of the random walk excursions one easily obtains the following expression for the constrained partition function:
916: \begin{equation} \label{eq:step1}
917: Z_{N,\go}^{\rc}\, = \,
918: \bE \left[
919: \prod_{j=1}^{\iota_N} \exp\big( \Psi^\go(\tau_{j-1}, \tau_j)\big);
920: \, \tau_{\iota_N}=N
921: \right],
922: \end{equation}
923: where $\iota_N = \sup\{k:\, \tau_k\le N\}$ and we have introduced
924: the \textsl{integrated Hamiltonian} $\Psi^\go (n_1,n_2)$, which gives the energetic contribution of an excursion from~$n_1$ to~$n_2$:
925: \begin{equation}
926: \label{eq:def_Psi}
927: \Psi^\go (n_1,n_2)=\begin{cases}
928:     \go ^{(0)}_{n_2} \,+\, \big( \tilde  \go ^{(0)}_{n_2} - \go^{(+1)}_{n_2} \big) & \text{if } n_2=n_1+1 \\
929:     \rule{0pt}{24pt} \go ^{(0)}_{n_2} \,+\, \displaystyle \log \bigg[ \frac 12 \Big( 1 + \exp \sum_{n=n_1+1}^{n_2} \big(\go_n ^{(-1)}-\go_n ^{(+1)}\big)
930: \Big) \bigg]  & \text{if } n_2>n_1 +1 \\
931:     \rule{0pt}{11pt} 0 & \text{otherwise}\,.
932: \end{cases}
933: \end{equation}
934: 
935: \smallskip
936: 
937: Now we are going to use in an essential way the fact that our charges are~$T$--periodic. In fact a look at \eqref{eq:def_Psi} shows that the energy $\Psi^\go (n_1,n_2)$ of an excursion from~$n_1$ to~$n_2$ is
938: a function only of $(n_2-n_1)$, $[n_1]$ and~$[n_2]$, where by~$[\,\cdot\,]$ we mean the equivalence class modulo~$T$, see \S~\ref{sec:order_par}. More precisely for $n_1 \in \ga$, $n_2 \in \gb$ and $\ell = n_2-n_1$ we have $\Psi^\go(n_1,n_2) = \Phi^\go_{\ga,\gb} (\ell)$, where~$\Phi^\go$ was defined in~\eqref{eq:def_Phi}. Then recalling the law~$K(n)$ of the first return, introduced in~\eqref{eq:first_ret}, we can rewrite \eqref{eq:step1} as
939: \begin{equation} \label{eq:zero_set_dec}
940: Z_{N,\go}^{\rc}\, = \, \sum_{k=1}^N \sumtwo{t_0,\ldots,t_k \in \N}{0=:t_0 < t_1 < \ldots < t_k:=N} \, \prod_{j=1}^k K\left( t_j - t_{j-1}\right) \; \exp\big(\Phi ^\go_{[t_{j-1}], [t_j]} (t_j - t_{j-1}) \big)\,.
941: \end{equation}
942: This decomposition of $Z_{N,\go}^{\rc}$ according to the random walk excursions makes explicit the link with Markov Renewal Theory. In fact using the kernel $M_{\ga,\gb}(x)$ introduced in \eqref{eq:matrix_cgz} we can rewrite it as
943: \begin{equation} \label{eq:conv1}
944: \begin{split}
945:     Z_{N,\go}^{\rc} &= \sum_{k=1}^N \sumtwo{t_0,\ldots,t_k \in \N}{0=:t_0 < t_1 < \ldots < t_k:=N} \, \prod_{j=1}^k M_{[t_{j-1}],[t_j]}(t_j-t_{j-1}) \\
946:     &= \sum_{k=1}^N \sumtwo{t_0,\ldots,t_k \in \N}{0=:t_0 < t_1 < \ldots < t_k:=N} \big[ M(t_1) \cdot M(t_2-t_{1}) \cdot \ldots \cdot M(N-t_{k-1}) \big]_{0,[N]} \\
947:     &= \sum_{k=0}^\infty \big[ M^{*k} \big] _{[0],[N]}(N)\,.
948: \end{split}
949: \end{equation}
950: Therefore it is natural to introduce the kernel $\cZ_{\ga,\gb}(x)$ defined by
951: \begin{align} \label{eq:expr_zeta1}
952:     \cZ_{\ga,\gb}(x) = \sum_{k=0}^\infty \big[ M^{*k} \big] _{\ga,\gb}(x)\,,
953: \end{align}
954: so that $Z_{N,\go}^{\rc} = \cZ_{[0],[N]}(N)$. More generally $\cZ_{\ga,\gb}(x)$ for $[x] = \gb - \ga$ can be interpreted as the partition function of a directed polymer of size~$x$ that starts at a site $(M,0)$, with $[M]=\ga$, and which is pinned at the site~$(M+x,0)$.
955: 
956: \smallskip
957: 
958: Our purpose is to get the precise asymptotic behavior of $\cZ_{\ga,\gb}(x)$ as~$x\to\infty$, from which we will obtain the asymptotic behavior of $Z_{N,\go}^{\rc}$ and hence the proof of Theorem~\ref{th:as_Z}.
959: It is clear that equation \eqref{eq:expr_zeta1} is the same as equation \eqref{eq:renewal_kernel}, except for the fact that in general the kernel~$M$ has no reason to be semi--Markov. Nevertheless we will see that with some transformations one can reduce the problem to a semi--Markov setting.
960: 
961: \smallskip
962: 
963: It turns out that for the derivation of the asymptotic behavior of~$\cZ_{\ga,\gb}(x)$ it is not necessary to use the specific form~\eqref{eq:matrix_cgz} of the kernel~$M_{\ga,\gb}(x)$, the computations being more transparent if carried out in a general setting. For these reasons, in the following we will assume that $M_{\ga,\gb}(x)$ is a generic $T$--periodic kernel such that the matrix~$B_{\ga,\gb}$ defined by~\eqref{eq:def_B} is finite. While these assumption are sufficient to yield the asymptotic behavior of~$\cZ_{\ga,\gb}(x)$ when $\gd^\go > 1$, for the cases $\gd^\go < 1$ and $\gd^\go = 1$ it is necessary to know the asymptotic behavior as~$x\to\infty$ of~$M_{\ga,\gb}(x)$ itself. Notice that our setting is an \textsl{heavy--tailed} one: more precisely we will assume that
964: for every $\ga,\gb\in\bbS$:
965: \begin{equation} \label{eq:asympt_cgz}
966: \exists \limtwo{x\to\infty}{[x]=\gb-\ga} x^{3/2}\, M_{\ga,\gb}(x) \;=:\; L_{\ga,\gb} \in (0,\infty) \,.
967: \end{equation}
968: From equation \eqref{eq:def_Phi} it is easy to check that the kernel $M_{\ga,\gb}(x)$ defined by \eqref{eq:matrix_cgz} does satisfy \eqref{eq:asympt_cgz} (see Section~\ref{sec:infvol} for more details on this issue).
969: 
970: \smallskip
971: 
972: For ease of exposition, we will treat separately the three cases $\gd^\go>1$, $\gd^\go <1$ and $\gd^\go = 1$.
973: 
974: \smallskip
975: \subsection{The localized regime $(\gd^\go > 1)$}
976: \label{sec:as_Z_loc}
977: 
978: The key idea is to introduce the following exponential perturbation of the kernel $M$ (cf. \cite[Theorem 4.6]{cf:Asm}), depending on the positive real parameter~$b$:
979: \begin{equation*}
980:     A^{b}_{\ga,\gb}(x) := M_{\ga,\gb}(x)\, e^{-bx}\,.
981: \end{equation*}
982: Let us denote by $\Delta(b)$ the Perron--Frobenius eigenvalue of the matrix $\sum_x A_{\ga,\gb}^{b}(x)$. As the entries of this matrix are analytic and nonincreasing functions of~$b$, $\Delta(b)$ is analytic and nonincreasing too, hence strictly decreasing because $\Delta(0) = \gd^\go > 1$ and $\Delta(\infty) = 0$. Therefore there exists a single value $\tf_\go>0$ such that $\Delta\big(\,\tf_\go\,\big)=1$, and we denote by $\{\zeta_\ga\}_\ga$, $\{\xi_\ga\}_\ga$ the Perron--Frobenius left and right eigenvectors of $\sum_x A_{\ga,\gb}^{\tf_\go}(x)$, chosen to have (strictly) positive components and normalized in such a way that $\sum_\ga \zeta_\ga\, \xi_\ga = 1$ (of course there is still a degree of freedom in the normalization, which however is immaterial).
983: 
984: \smallskip
985: 
986: Now we set
987: \begin{equation} \label{eq:def_Gamma>}
988:     \Gamma^>_{\ga,\gb} (x) := A^{\tf_\go}_{\ga,\gb} (x) \, \frac{\xi_\gb}{\xi_\ga} = M_{\ga,\gb} (x)\, e^{-\tf_\go x}\, \frac{\xi_\gb}{\xi_\ga}\,,
989: \end{equation}
990: and it is immediate to check that $\Gamma^>$ is a semi--Markov kernel. Furthermore, we can rewrite~\eqref{eq:expr_zeta1} as
991: \begin{equation} \label{eq:Z_loc}
992:     \cZ_{\ga,\gb}(x) := e^{\tf_\go x}\, \frac{\xi_\ga}{\xi_\gb}\, \sum_{k=0}^\infty \big[ (\Gamma^>)^{*k} \big] _{\ga,\gb}(x) = e^{\tf_\go x}\, \frac{\xi_\ga}{\xi_\gb}\, \cU_{\ga,\gb}(x) \,,
993: \end{equation}
994: where $\cU_{\ga,\gb}(x)$ is nothing but the Markov--Green function associated to the semi--Markov kernel~$\Gamma^>_{\ga,\gb}(x)$. Therefore the asymptotic behavior of $\cZ_{\ga,\gb}(x)$ is easily obtained applying Blackwell's Renewal Theorem~\eqref{eq:per_renewal_theorem}. To this end, let us compute the mean $\mu$ of the semi--Markov kernel $\Gamma^>$: it is easily seen that the invariant measure of the associated modulating chain is given by $\{\zeta_\ga \xi_\ga\}_\ga$, therefore
995: \begin{align*}
996:     \mu & = \sum_{\ga,\gb \in \bbS} \sum_{x\in\N} x \, \zeta_\ga \, \xi_\ga \, \Gamma_{\ga,\gb}^>(x) = \sum_{\ga,\gb \in \bbS} \sum_{x\in\N} x e^{-\tf_\go x} \, \zeta_\ga\, M_{\ga,\gb}(x)\, \xi_\gb\\
997:     & = - \bigg( \frac{\partial}{\partial b} \Delta(b) \bigg) \bigg|_{b = \tf_\go} \ \ \in \ (0,\infty)\,,
998: \end{align*}
999: (for the last equality see for example \cite[Lemma 2.1]{cf:BG}). Coming back to \eqref{eq:Z_loc}, we can now apply Blackwell's Renewal Theorem \eqref{eq:per_renewal_theorem} obtaining the desired asymptotic behavior:
1000: \begin{equation}\label{eq:as_Z_loc}
1001:     \cZ_{\ga,\gb}(x) \; \sim \; \xi_\ga \, \zeta_\gb \,  \frac{T}{\mu} \,
1002: \exp \left( \tf_\go\, x \right) \qquad \quad x\to\infty,\quad [x]=\gb-\ga\,.
1003: \end{equation}
1004: In particular, for $\ga=[0]$ and $\gb=\eta$ we have part~(1) of Theorem~\ref{th:as_Z},
1005: where $C^>_{\go,\eta} = \xi_0 \zeta_\eta T /\mu$.
1006: 
1007: \smallskip
1008: \subsection{The strictly delocalized case $(\gd^\go < 1)$}
1009: \label{sec:as_Z_del}
1010: 
1011: We prove that the asymptotic behavior of~$\cZ_{\ga,\gb}(x)$ when $\gd^\go < 1$ is given by
1012: \begin{equation}\label{eq:as_Z_del}
1013:     \cZ_{\ga,\gb}(x) \; \sim \; \Big( \big[(1-B)^{-1} L \, (1-B)^{-1} \big]_{\ga,\gb} \Big) \: \frac{1}{x^{3/2}} \qquad \quad x\to\infty,\quad [x]=\gb-\ga \,,
1014: \end{equation}
1015: where the matrixes $L$ and $B$ have been defined in \eqref{eq:asympt_cgz} and \eqref{eq:def_B}.
1016: In particular, taking $\ga=[0]$ and $\gb=\eta$, \eqref{eq:as_Z_del} proves part~(2) of
1017: Theorem~\ref{th:as_Z} with
1018: \begin{equation*}
1019:     C^<_{\go,\eta} := \big[(1-B)^{-1} L \, (1-B)^{-1} \big]_{0,\eta} \,.
1020: \end{equation*}
1021: 
1022: \smallskip
1023: 
1024: To start with, we prove by induction that for every $n\in\N$
1025: \begin{equation} \label{eq:sum}
1026: \sum_{x\in\N} [M^{*n}]_{\ga,\gb} (x) = [B^n]_{\ga,\gb}\,.
1027: \end{equation}
1028: The $n=1$ case is the definition of~$B$, while for $n\geq 1$
1029: \begin{align*}
1030: \sum_{x\in\N} M^{*(n+1)}(x) & \;=\; \sum_{x\in\N} \sum_{z\le x} M^{*n}(z) \cdot M(x-z) \;=\; \sum_{z\in\N}\, M^{*n}(z) \cdot \sum_{x\ge z}\, M(x-z) \\
1031: & \;=\; \sum_{z\in\N}\, M^{*n}(z) \cdot B \;=\; B^n \cdot B \;=\; B^{n+1} \,.
1032: \end{align*}
1033: 
1034: \smallskip
1035: 
1036: Next we claim that, if \eqref{eq:asympt_cgz} holds, then for every $\ga,\gb \in \bbS$
1037: \begin{equation} \label{eq:as_M_k}
1038: \exists \limtwo{x\to\infty}{[x]=\gb-\ga} \, x^{3/2} \big[M^{*k}\big]_{\ga,\gb}(x) \;=\; \sum_{i=0}^{k-1} \big[B^i\cdot L \cdot B^{(k-1)-i}\big]_{\ga,\gb}\,.
1039: \end{equation}
1040: We proceed by induction on~$k$. The $k=1$ case is given by~\eqref{eq:asympt_cgz}, and we have that
1041: \[
1042: M^{*(n+1)}(x) = \sum_{y=1}^{x/2} \bigg( M(y)\cdot M^{*n}(x-y) \;+\; M(x-y) \cdot M^{*n}(y) \bigg)
1043: \]
1044: (strictly speaking this formula is true only when $x$ is even, however the odd~$x$ case is analogous). By the inductive hypothesis equation \eqref{eq:as_M_k} holds for every $k\leq n$, and in particular this implies that $\{x^{3/2} [M^{*k}]_{\ga,\gb}(x)\}_{x\in\N}$ is a bounded sequence. Therefore we can apply Dominated Convergence and \eqref{eq:sum}, getting
1045: \begin{align*}
1046:     \exists \limtwo{x\to\infty}{[x]=\gb-\ga} & \, x^{3/2} \big[M^{*(n+1)}\big]_{\ga,\gb}(x) \\
1047:     & \;=\; \sum_\gamma \sum_{y=1}^{\infty} \bigg( M_{\ga,\gamma}(y)\, \sum_{i=0}^{n-1} \big[B^i\cdot L \cdot B^{(n-1)-i}\big]_{\gamma,\gb} \;+\; L_{\ga,\gamma} \big[M^{*n}\big]_{\gamma,\gb}(y) \bigg)\\
1048: & \;=\; \sum_\gamma  \bigg( B_{\ga,\gamma}\, \sum_{i=0}^{n-1} \big[B^i\cdot L \cdot B^{(n-1)-i}\big]_{\gamma,\gb} \;+\; L_{\ga,\gamma} \big[ B^{*n} \big]_{\gamma,\gb} \bigg)\\
1049: & \;=\; \sum_{i=0}^{n} \big[B^i\cdot L \cdot B^{n-i}\big]_{\ga,\gb} \,.
1050: \end{align*}
1051: 
1052: \smallskip
1053: 
1054: Our purpose is to apply the asymptotic result \eqref{eq:as_M_k} to the terms of \eqref{eq:expr_zeta1}, hence we need a bound to apply Dominated Convergence. What we are going to show is that
1055: \begin{equation}\label{eq:dom_con}
1056:     x^{3/2} \, \big[M^{*k}\big]_{\ga,\gb}(x) \;\le\; C\, k^3\, \big[ B^k \big]_{\ga,\gb}
1057: \end{equation}
1058: for some positive constant $C$ and for all $\ga,\gb\in\bbS$ and $x,k\in\N$. Observe that the r.h.s. above, as a function of~$k$, is a summable sequence because the matrix~$B$ has spectral radius $\gd^\go<1$. We proceed again by induction: for the $k=1$ case, thanks to \eqref{eq:asympt_cgz}, it is possible to find~$C$ such that \eqref{eq:dom_con} holds true (this fixes $C$ once for all). Now assuming that \eqref{eq:dom_con} holds for all $k < n$ we show that it does also for~$k=n$ (we suppose for simplicity that $n=2m$ is even, the odd~$n$ case being analogous). Then we have (assuming that also~$x$ is even for simplicity)
1059: \begin{align*}
1060:     x^{3/2} \, \big[M^{*2m}\big]_{\ga,\gb}(x) &\;=\; 2 \sum_{y=1}^{x/2} \sum_{\gamma \in \bbS} \big[M^{*m}\big]_{\ga,\gamma}(y)\; x^{3/2} \big[M^{*m}\big]_{\gamma,\gb}(x-y)\\
1061:     &\;\le\; 2 \cdot 2^{3/2} \, C \, m^{3} \sum_{y=1}^{x/2} \sum_{\gamma \in \bbS} \big[M^{*m}\big]_{\ga,\gamma}(y)\; \big[B^{m}\big]_{\gamma,\gb} \\
1062:     &\;\le\; C\, (2m)^3 \big[ B^{2m} \big]_{\ga,\gb}\,,
1063: \end{align*}
1064: where we have applied \eqref{eq:sum}, and \eqref{eq:dom_con} is proven.
1065: 
1066: \smallskip
1067: 
1068: We can finally obtain the asymptotic behavior of~$\cZ_{\ga,\gb}(x)$ applying the bound \eqref{eq:as_M_k} to \eqref{eq:expr_zeta1}, using Dominated Convergence thanks to \eqref{eq:dom_con}. In this way we get
1069: \begin{align*}
1070:     \exists \limtwo{x\to\infty}{[x]=\gb-\ga} & \; x^{3/2} \cZ_{\ga,\gb}(x) \;=\; \sum_{k=1}^\infty \sum_{i=0}^{k-1} \big[B^i\cdot L \cdot B^{(k-1)-i}\big]_{\ga,\gb} \\
1071:     &\;=\; \sum_{i=0}^\infty \sum_{k=i+1}^\infty \big[B^i\cdot L \cdot B^{(k-1)-i}\big]_{\ga,\gb} \;=\; \sum_{i=0}^\infty \big[B^i\cdot L \cdot (1-B)^{-1}\big]_{\ga,\gb} \\
1072:     &\;=\; \big[(1-B)^{-1}\cdot L \cdot (1-B)^{-1}\big]_{\ga,\gb}\,,
1073: \end{align*}
1074: and equation \eqref{eq:as_Z_del} is proven.
1075: 
1076: \smallskip
1077: 
1078: \subsection{The critical case $(\gd^\go = 1)$}
1079: \label{sec:as_Z_cri}
1080: 
1081: In the critical case the matrix $B$ defined in~\eqref{eq:def_B} has Perron--Frobenius eigenvalue equal to~1. Let $\{\zeta_\ga\}_\ga$, $\{\xi_\ga\}_\ga$ denote its corresponding left and right eigenvectors, always chosen to have positive components and normalized so that $\sum_\ga \zeta_\ga\, \xi_\ga = 1$. Then it is immediate to check that the kernel
1082: \begin{equation}\label{eq:def_Gamma=}
1083:     \Gamma^=_{\ga,\gb}(x) \;:=\; M_{\ga,\gb} (x) \, \frac{\xi_\gb}{\xi_\ga}
1084: \end{equation}
1085: is semi--Markov, and the corresponding Markov--Green function $U_{\ga,\gb}(x)$ is given by
1086: \begin{equation} \label{eq:def_U}
1087:     U_{\ga,\gb}(x) \;:=\; \sum_{k=0}^\infty \big[ (\Gamma^=)^{*k} \big]_{\ga,\gb} (x) \;=\; \frac{\xi_\gb}{\xi_\ga} \, \cZ_{\ga,\gb}(x)\,,
1088: \end{equation}
1089: where the last equality follows easily from~\eqref{eq:expr_zeta1}. We are going to derive the asymptotic behavior of~$U_{\ga,\gb}(x)$, and from the above relation we will get the analogous result for~$\cZ_{\ga,\gb}(x)$.
1090: 
1091: \smallskip
1092: 
1093: Denoting by $\{(T_k,J_k)\}$ under~$\bbP$ the Markov--renewal process generated by the semi--Markov kernel $\Gamma^=_{\ga,\gb}(x)$, for $U_{\ga,\gb}(x)$ we have the probabilistic interpretation \eqref{eq:prob_interpret}, that we rewrite for convenience
1094: \begin{equation} \label{eq:prob_interpret2}
1095:     U_{\ga,\gb} (x) = \bbP_\ga \big[ \exists k \ge 0 :\; T_0+\ldots +T_k=x\;,\; J_{k} = \gb \big]\,.
1096: \end{equation}
1097: For $\gb \in \bbS$ we introduce the sequence of stopping times~$\{\kappa^{(\gb)}_n\}_{n \geq 0}$
1098: corresponding to the visit of the chain~$\{J_k\}$ to the state~$\gb$:
1099: \begin{equation}\label{defkappa}
1100:     \kappa_0^{(\gb)} := \inf \{k \geq 0:\ J_k = \gb\} \qquad \quad \kappa^{(\gb)}_{n+1}
1101: := \inf \{k > \kappa^{(\gb)}_n:\ J_k = \gb\}\,,
1102: \end{equation}
1103: and we define the process $\{T^{(\gb)}_n\}_{n \geq 0}$ by setting
1104: \begin{equation}\label{defTkappa}
1105:     T_0^{(\gb)} := T_0 + \ldots + T_{\kappa^{(\gb)}_0} \qquad \quad T_{n}^{(\gb)} :T_{\kappa^{(\gb)}_{n-1} + 1} + \ldots + T_{\kappa^{(\gb)}_{n}} \,.
1106: \end{equation}
1107: The key point is that under~$\bbP_\ga$ the random variables $\{T^{(\gb)}_n\}$ are the interarrival times of a (possibly delayed) \textsl{classical renewal process}, equivalently the sequence $\{T^{(\gb)}_n\}_{n \geq 1}$ is IID and independent of~$T_0^{(\gb)}$. We denote for $x\in\N$ by $q^{(\gb)}(x)$ the (mass function of the) law of~$T_n^{(\gb)}$ for $n\geq 1$, while the law of~$T_0^{(\gb)}$ under~$\bbP_\ga$ is denoted by~$q^{(\ga;\gb)}(x)$. Since clearly
1108: \begin{equation*}
1109:     \big\{ \exists k \ge 0 :\; T_0+\ldots +T_k=x\;,\; J_{k} = \gb \big\} \quad \iff \quad \big\{ \exists n \ge 0 :\; T^{(\gb)}_0+\ldots +T^{(\gb)}_n=x \big\}\,,
1110: \end{equation*}
1111: from \eqref{eq:prob_interpret2} we get
1112: \begin{equation} \label{eq:cl_ren_pr}
1113:     U_{\ga,\gb} (x) = \bbP_\ga \big[ \exists n \ge 0 :\; T^{(\gb)}_0+\ldots +T^{(\gb)}_n=x  \big] = \bigg( q^{(\ga;\gb)} * \sum_{n=0}^\infty \big( q^{(\gb)} \big) ^{*n} \bigg) (x) \,,
1114: \end{equation}
1115: which shows that $U_{\ga,\gb}(x)$ is indeed the Green function of the classical renewal process whose interarrival times are the $\{T_n^{(\gb)}\}_{n\ge 0}$.
1116: 
1117: \smallskip
1118: 
1119: Now we claim that the asymptotic behavior of $q^{(\gb)}(x)$ as $x\to\infty$, $x\in\gb$, is given by
1120: \begin{equation} \label{eq:cb1}
1121:     q^{(\gb)}(x) \; \sim \; \frac{c_\gb}{x^{3/2}} \qquad \quad
1122: c_\gb \;:=\; \frac1{\zeta_\gb\, \xi_\gb}
1123: \, \sum_{\ga,\gamma} \zeta_\ga \, L_{\ga,\gamma}\, \xi_\gamma  \;>\;0\,,
1124: \end{equation}
1125: see \S~\ref{app:qb} for a proof of this relation. Then the asymptotic behavior of \eqref{eq:cl_ren_pr} is given by
1126: \begin{equation}\label{eq:doney}
1127:     U_{\ga,\gb} (x) \;\sim\; \frac{T^2}{2\pi\,c_\gb} \, \frac{1}{\sqrt{x}} \qquad \quad x\to\infty,\quad [x] = \gb-\ga\,,
1128: \end{equation}
1129: as it follows by~\cite[Th.~B]{cf:Don97} (the factor~$T^2$ is due to our periodic setting). Combining equations \eqref{eq:def_U}, \eqref{eq:cb1} and \eqref{eq:doney} we finally get the asymptotic behavior of~$\cZ_{\ga,\gb}(x)$:
1130: \begin{equation}\label{eq:as_Z_cri}
1131:     \cZ_{\ga,\gb}(x) \;\sim\; \frac{T^2}{2\pi} \, \frac{\xi_\ga\, \zeta_\gb}
1132: { \sum_{\gamma,\gamma'} \zeta_\gamma \, L_{\gamma,\gamma'}\, \xi_{\gamma'} }
1133: \: \frac{1}{\sqrt{x}} \qquad \quad x\to\infty,\quad [x]=\gb-\ga \,.
1134: \end{equation}
1135: Taking $\ga=[0]$ and $\gb=\eta$, we have the proof of part~(3) of Theorem~\ref{th:as_Z}.
1136: 
1137: 
1138: 
1139: 
1140: 
1141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1142: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1143: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1144: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1145: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1146: 
1147: 
1148: 
1149: 
1150: \smallskip
1151: \section{Thermodynamic limits}
1152: \label{sec:infvol}
1153: 
1154: In this section we study the limit as $N\to\infty$ of the polymer
1155: measure $\bP^a_{N, \go}$, using the sharp asymptotics for the partition
1156: function obtained in the previous section.
1157: We recall that  $\bP^\rc_{N,\go}$ is a probability
1158: measure on $\Z^\N$, which we endow with the product topology.
1159: In particular, weak convergence on $\Z^\N$ means convergence
1160: of all finite dimensional marginals.
1161: 
1162: \medskip
1163: 
1164: We start giving a very useful decomposition of $\bP^a_{N,\go}$.
1165: The intuitive idea is that a path $(S_n)_{n\le N}$ can be split
1166: into two main ingredients:
1167: \smallskip
1168: \begin{itemize}
1169: \item the family $(\tau_k)_{k=0,1,\ldots}$ of \textsl{returns to zero}
1170: of $S$ (defined in \S~\ref{sec:RWexcurs});
1171: \item \rule{0pt}{13pt}the family of \textsl{excursions from zero}
1172: $(S_{i+\tau_{k-1}}:0\le i\le \tau_{k}-\tau_{k-1})_{k=1, 2, \ldots}$
1173: \end{itemize}
1174: \smallskip
1175: Moreover, since each excursion can be
1176: either positive or negative, it is also useful to consider separately
1177: the signs of the excursions $\sigma_k := \sign (S_{\tau_{k-1} + 1})$ and the
1178: absolute values $(e_k(i):=|S_{i+\tau_{k-1}}|:\, i=1,\ldots,\tau_{k}-\tau_{k-1})$.
1179: Observe that these are trivial for an
1180: excursion with length~$1$: in fact if $\tau_{k}=\tau_{k-1} + 1$
1181: then $\gs_k =0$ and~$e_k(0)=e_k(1)=0$.
1182: 
1183: \smallskip
1184: 
1185: Let us first consider the returns $(\tau_k)_{k\le\iota_N}$ under~$\bP_{N,\go}^a$,
1186: where $\iota_N = \sup\{k:\, \tau_k\le N\}$.
1187: The law of this process can be viewed as a probability measure $p^a_{N,\go}$
1188: on the class $\cA_N$ of subsets of $\{1,\ldots, N\}$: indeed
1189: for $A\in\cA_N$, writing
1190: %
1191: \begin{equation}\label{notA}
1192: A  = \{t_1,\ldots,t_{|A|}\}, \qquad
1193: 0 \, =: \, t_0<t_1<\cdots<t_{|A|} \, \leq \, N,
1194: \end{equation}
1195: %
1196: we can set
1197: %
1198: \begin{equation}\label{defp^a}
1199: p^a_{N,\go}(A) \, := \, \bP^a_{N,\go}(\tau_i= t_i, \ i\leq\iota_N).
1200: \end{equation}
1201: The measure $p^a_{N,\go}$ describes the zero set of the polymer of size~$N$,
1202: and it is analyzed in detail below. From the inclusion of $\cA_N$ into $\{0,1\}^\N$,
1203: the family of all subsets of~$\N$, $p_{N,\go}^a$ can be viewed as a measure
1204: on~$\{0,1\}^\N$ (this observation will be useful in the following).
1205: %
1206: 
1207: \smallskip
1208: 
1209: Now we pass to the signs: we can see that, given $(\tau_j)_{j\leq\iota_N}$,
1210: under $\bP^a_{N,\go}$ the signs $(\sigma_k)_{k\leq\iota_N}$ form
1211: an independent family. Conditionally on~$(\tau_j)_{j\leq\iota_N}$, the law of~$\gs_k$ is specified by:
1212: \begin{itemize}
1213: \item[-] if~$\tau_{k} = 1+\tau_{k-1} $, then $\gs_k = 0$;
1214: \item[-] \rule{0pt}{12pt}if~$\tau_{k} > 1+\tau_{k-1} $, then $\gs_k$ can take the two values~$\pm 1$ with
1215: \begin{equation}\label{defsigma}
1216: \bP^a_{N,\go}\Big(\gs_k=+1 \, \Big| \ (\tau_j)_{j\leq\iota_N}\Big)
1217: \, = \, \frac{1}
1218: {1 + \exp\left\{ -(\tau_{k} - \tau_{k-1}) \, h_\go + \Sigma_{[\tau_{k-1}],[\tau_{k}]} \right\}}\,.
1219: \end{equation}
1220: %where for the sake of this equation we set~$\tau_{\iota_N + 1} := N$.
1221: \end{itemize}
1222: Observe that when~$\tau_{\iota_N} < N$ (which can happen only for~$a=\rf$)
1223: there is a last (incomplete) excursion in the interval
1224: $\{0, \ldots, N\}$, and the sign of this excursion is also expressed by~\eqref{defsigma} for~$k=\iota_{N+1}$,
1225: provided we set~$\tau_{\iota_{N} + 1}:= N$.
1226: 
1227: 
1228: \smallskip
1229: 
1230: Finally we have the moduli: again, once $(\tau_{k-1},\sigma_k)_{1\leq k\leq\iota_N+1}$ are given,
1231: the excursions $(e_k)_{k=1,\dots,\iota_N+1}$ form an independent family. The conditional law of~$e_k(\cdot)$
1232: on the event $\{\tau_{k-1}=\ell_0, \ \tau_{k}=\ell_1\}$ and for $f=(f_i)_{i=1, \ldots, \ell_1 - \ell_0}$
1233: is, for $k\leq\iota_N$, given by
1234: %
1235: \begin{align}\label{defexc}
1236: \begin{split}
1237: & \bP^a_{N,\go}\Big( e_k(\cdot)=f \ \Big| \ (\tau_{j-1},\sigma_j)_{1\leq j\leq\iota_N+1} \Big) \\
1238: & = \ \bP\Big(S_i=f_i: \ i=1,\ldots,\ell_1-\ell_0 \ \Big| \
1239: S_i > 0: \ i=1, \ldots, \ell_1-\ell_0-1, \ S_{\ell_1-\ell_0}=0\Big)\,.
1240: \end{split}
1241: \end{align}
1242: %
1243: In the case~$\tau_{\iota_N} < N$ we have a last excursion
1244: $e_{\iota_N + 1}(\cdot)$: its conditional law, on the event $\{\tau_{\iota_N}=\ell<N\}$
1245: and for $f=(f_i)_{i=1, \ldots, N - \ell}$, is given by
1246: %
1247: \begin{align}\label{defexcl}
1248: \begin{split}
1249: &\bP^a_{N,\go}\Big( e_{\iota_N + 1}(\cdot)=f \ \Big| \ (\tau_{j-1},\sigma_j)_{1\leq j\leq\iota_N+1}
1250: \Big)\\
1251: & = \ \bP\Big(S_i=f_i: \ i=1,\ldots,N-\ell \ \Big| \
1252: S_i> 0: \ i = 1, \ldots, N-\ell\Big),
1253: \end{split}
1254: \end{align}
1255: %
1256: 
1257: \smallskip
1258: 
1259: We would like to stress that the above relations fully characterize
1260: the polymer measure $\bP^a_{N,\go}$. A remarkable fact is that,
1261: conditionally on $(\tau_k)_{k\in \N}$, the joint distribution
1262: of $(\gs_j,e_j)_{j\leq \iota_N}$ \textsl{does not depend on $N$}: in this sense,
1263: all the $N$--dependence is contained in the measure $p^a_{N,\go}$.
1264: 
1265: For this reason, this section is mainly devoted to the study of the
1266: asymptotic behavior of the zero set measures~$p^a_{N,\go}$ as $N\to\infty$.
1267: The main result is that $p^\rc_{N,\go}$ and $p^\rf_{N,\go}$  have the
1268: same weak limit $p_\go$ on~$\{0,1\}^\N$ as~$N\to\infty$ (with some
1269: restrictions when~$\go \in \cP^<$).
1270: Once this is proven, it follows easily that also the polymer
1271: measure~$\bP_{N,\go}^a$ converges
1272: to a limit measure $\bP_\go$ on~$\Z^\N$, constructed by pasting
1273: the excursion over the limit zero set. More precisely, $\bP_\go$
1274: is the measure under which the processes $(\tau_j)$, $(\sigma_j)$
1275: and~$(e_j)$ have the following laws:
1276: \begin{itemize}
1277: \smallskip
1278: \item the law of the~$(\tau_j)_{j\in\N}$ is determined in an obvious way by the
1279: limiting zero set measure~$p_\go$;
1280: \smallskip
1281: \item then, conditionally on the~$(\tau_j)_{j\in\N}$, the process $(\sigma_j)_{j\in\N}$ is an independent
1282: one with marginal laws given by~\eqref{defsigma};
1283: \smallskip
1284: \item finally, conditionally on~$(\tau_j,\gs_j)_{j\in\N}$, on the event $\{\tau_{k-1}=\ell_0, \tau_{k}=\ell_1\}$
1285: with $\ell_0<\ell_1<\infty$ the law of $e_k$ is given by the r.h.s. of (\ref{defexc}).
1286: We have to consider also the case $\ell_0 < \infty$, $\ell_1 = \infty$, because
1287: in the regime $\delta^\go<1$ it turns out that $\bP_\go(\tau_{k}=\infty) > 0$ (see
1288: below and \S~\ref{app:ther}): in this case
1289: the law of $e_k$ is given for any~$n\in\N$ and for $f=(f_i)_{i=1, \ldots, n}$ by:
1290: %
1291: \begin{align}\label{defexclul}
1292: \begin{split}
1293: &\bP_{\go}\Big(e_{k}(i)=f_i\,:\ i = 1, \ldots, n \ \Big| \ (\tau_j,\gs_j)_{j\in\N} \Big) \, = \,
1294: \bP^+\Big(S_i=f_i\,:\ i = 1, \ldots, n\Big)
1295: \\ & \qquad := \ \lim_{N\to\infty} \bP\Big(S_i=f_i\,: \ i=1,\ldots,n \ \Big| \
1296: S_i> 0: \ i = 1, \ldots, N\Big),
1297: \end{split}
1298: \end{align}
1299: %
1300: where the existence of such limit is well known: see e.g. \cite{cf:G}.
1301: \end{itemize}
1302: 
1303: \smallskip
1304: \subsection{Law of the zero level set in the free and constrained cases}
1305: \label{sec:zero_set}
1306: Let us describe more explicitly $p^a_{N,\go}(A)$, using the (strong)
1307: Markov property of $\bP^a_{N,\go}$. We use throughout the chapter the notation
1308: (\ref{notA}). Recalling the definition \eqref{eq:matrix_cgz}
1309: of $M_{\ga,\gb}(t)$, we have:
1310: %
1311: \begin{itemize}
1312: %
1313: \item for $a= \rc$ and $A\in\cA_N$: \ $p^\rc_{N,\go}(A)\ne 0$ \ if and
1314: only if \  $t_{|A|}=N$, \ and in this case:
1315: \[
1316:  p^\rc_{N,\go}(A) \, = \,
1317: \frac 1{Z^\rc_{N,\go}} \, \prod_{i=1}^{|A|}
1318:  M_{[t_{i-1}],[t_i]}(t_i-t_{i-1})
1319: \]
1320: %
1321: \item \rule{0pt}{13pt}for $a= \rf$ and $A\in\cA_N$:
1322: \[
1323:  p^\rf_{N,\go}(A) \, = \,
1324: \frac 1{Z^\rf_{N,\go}} \, \left[\prod_{i=1}^{|A|}
1325:  M_{[t_{i-1}],[t_i]}(t_i-t_{i-1})\right] P(N-t_{|A|}) \,
1326:  \exp \Big( \tPhi_{[t_{|A|}],[N]}(N-t_{|A|}) \Big).
1327: \]
1328: where $P(n) := \sum_{k=n+1}^\infty K(k) = \sum_{k=n+1}^\infty \bP (\tau_1 = k)$ and we have introduced
1329: \begin{equation} \label{eq:def_tPhi}
1330:     \tPhi_{\ga,\gb}(\ell) \;:=\; \log \bigg[ \frac 12 \Big( 1 + \exp \big( -\ell h_\go + \Sigma_{\ga,\gb} \big) \Big) \bigg] \; \ind_{(\ell\, > 1)} \; \ind_{(\ell \, \in \gb-\ga)}\,,
1331: \end{equation}
1332: which differs from~$\Phi$ in not having the terms of interaction with the interface, cf.~\eqref{eq:def_Phi}.
1333: %
1334: \end{itemize}
1335: %
1336: 
1337: \smallskip
1338: 
1339: We are going to show that, for any value of~$\gd^\go$, the
1340: measure~$p^a_{N,\go}$ on~$\{0,1\}^\N$ converges as~$N\to\infty$
1341: (with some restrictions if~$\go \in \cP^<$) to a limit measure
1342: under which the process $([\tau_k],\tau_k-\tau_{k-1})_{k\in\bbN}$
1343: is a \textsl{Markov renewal process}. Moreover, we will compute
1344: explicitly the corresponding semi--Markov kernel, showing that
1345: the returns to zero are
1346: \begin{enumerate}
1347: %
1348: \item integrable if $\delta^\go>1$ (localized regime);
1349: %
1350: \item \rule{0pt}{13pt}defective if $\delta^\go<1$ (strictly delocalized regime);
1351: %
1352: \item \rule{0pt}{13pt}non integrable if $\delta^\go=1$ (critical regime).
1353: %
1354: \end{enumerate}
1355: Thanks to the preceding observations, this will complete the proof of~Theorem~\ref{th:infvol}. We stress that the key result in our derivation is given by the sharp asymptotics of the partition function $Z^\rc_{N,\go}$ obtained in the previous section.
1356: 
1357: \smallskip
1358: 
1359: Before going into the proof, we give some preliminary material which is useful for all values of~$\delta^\go$.
1360: For $k\in\N$ we define the shift operator:
1361: \[
1362: \theta_k:\R^\N\mapsto\R^\N, \qquad
1363: \theta_k\zeta := \zeta_{[k+\cdot]},
1364: \]
1365: and it is easy to check that the following relations hold true:
1366: %
1367: \begin{equation}\label{eq:ZcZ}
1368: Z_{N-k,\theta_k\go}^{\rc} = \cZ_{[k],[N]}(N-k), \qquad k\leq N.
1369: \end{equation}
1370: %
1371: %
1372: \begin{equation}\label{eq:relfc}
1373: Z_{N,\go}^{\rf} \, = \, \sum_{t=0}^N Z_{t,\go}^{\rc} \, P(N-t) \,
1374: \exp\left(\tPhi_{[t],[N]}(N-t) \right),
1375: \end{equation}
1376: %
1377: %
1378: \begin{equation}\label{eq:relpmz}
1379: \bP^a_{N, \go} \left( \tau_1 =k \right)
1380: \, = \, M_{0,[k]}(k) \ \frac{Z^a_{N-k,\theta_k\go}}{Z^a_{N,\go}},
1381: \qquad 1\leq k\leq N, \quad a=\rc,\rf.
1382: \end{equation}
1383: %
1384: Finally, using \eqref{eq:as_K}, \eqref{eq:matrix_cgz} and~\eqref{eq:def_Phi} it is easy to see that
1385: \eqref{eq:asympt_cgz} holds true, namely
1386: %
1387: \begin{equation}\label{eq:as_L_div}
1388: \exists \limtwo{x\to\infty}{[x]=\gb-\ga} \, x^{3/2} \, M_{\ga,\gb}(x)
1389: \;=\;  L_{\ga, \gb},
1390: \end{equation}
1391: %
1392: where:
1393: %
1394: \begin{equation}\label{defL}
1395: L_{\ga, \gb} \, = \,
1396: %
1397: \begin{cases}
1398: \displaystyle  c_K \, \frac 12 \, \Big(1+
1399: \exp\big( \Sigma_{\ga,\gb}\big)\Big) \exp(\go ^{(0)}_{\gb})
1400: & \text{if\ \ } h_\go = 0\\
1401: \displaystyle\rule{0pt}{22pt} c_K \, \frac 12 \, \exp(\go ^{(0)}_{\gb})
1402: & \text{if\ \ } h_\go > 0
1403: \end{cases}
1404: %
1405: \; .
1406: \end{equation}
1407: %
1408: Since also the asymptotic behavior of $P(\ell) \exp(\tPhi_{\ga,\gb}(\ell))$
1409: will be needed, we set
1410: %
1411: \begin{equation} \label{eq:as_tPhi}
1412:     \tL_{\ga,\gb} \;:=\; \lim_{\ell \to \infty,\ \ell \in \gb-\ga} \sqrt{\ell}
1413: \, P(\ell) \, e^{\tPhi_{\ga,\gb}(\ell)} \;=\;
1414: \begin{cases}
1415: \displaystyle  c_K \big( 1 + \exp(\Sigma_{\ga,\gb}) \big) & \text{if\ \ } h_\go = 0\\
1416: \displaystyle\rule{0pt}{22pt} c_K & \text{if\ \ } h_\go > 0
1417: \end{cases}
1418: \;,
1419: \end{equation}
1420: %
1421: as it follows easily from~\eqref{eq:def_tPhi} and from the fact
1422: that~$P(\ell) \sim 2 \, c_K / \sqrt{\ell}$ as~$\ell\to\infty$.
1423: 
1424: 
1425: %We point out the following fact, which will be important in
1426: %Remark \ref{sit2} and in the proof of Proposition \ref{pr:infvol2}:
1427: %
1428: %\begin{rem}\label{sit}{\rm
1429: %$(L_{\ga, \gb})_{\ga, \gb\in\bbS}$ and $(\tL_{\ga, \gb})_{\ga, \gb\in\bbS}$
1430: %are constant in $\ga$ if either $h_\go>0$ or $h_\go=0$ and $\Sigma_{\ga,\gb}=0$ \
1431: %$\forall \ga,\gb$.
1432: %}\end{rem}
1433: %
1434: 
1435: \smallskip
1436: \subsection{The localized regime $(\delta^\go>1)$}
1437: 
1438: We prove point (1) of Theorem \ref{th:infvol}. More precisely, we
1439: prove the following:
1440: %
1441: \begin{proposition}\label{pr:infvol1}
1442: If $\delta^\go>1$ then
1443: the polymer measures $\bP^\rf_{N, \go}$ and $\bP^\rc_{N, \go}$
1444: converge as $N\to\infty$ to the same limit
1445: $\bP_\go$, under which $([\tau_k],\tau_k-\tau_{k-1})_{k\in\bbN}$
1446: is a Markov renewal process with semi-Markov kernel
1447: $(\Gamma^>_{\ga,\gb}(x): \ga,\gb\in\bbS, x\in\N)$.
1448: \end{proposition}
1449: %
1450: \noindent
1451: For the definition of $\Gamma^>$ see (\ref{eq:def_Gamma>}).
1452: 
1453: 
1454: \smallskip
1455: \subsubsection{Proof of Proposition \ref{pr:infvol1}.}
1456: We prove first the case $a=\rc$.
1457: By \eqref{eq:ZcZ}, (\ref{eq:relpmz})
1458: and by the asymptotics of $\cZ$ in (\ref{eq:as_Z_loc}) above,
1459: we have for all $\ga,\gb,\gga\in\bbS$ and $\ell\in\ga$, $m\in\gb$
1460: \[
1461: \exists \limtwo{N\to\infty}{N\in\gga} \
1462: \frac{Z^\rc_{N-m,\theta_m\go}}{Z^\rc_{N-\ell,\theta_\ell\go}}
1463: \, = \, \limtwo{N\to\infty}{N\in\gga} \
1464: \frac{\cZ_{\gb,\gga}(N-m)}{\cZ_{\ga,\gga}(N-\ell)}
1465: \, = \, e^{-\tf_\go k} \, \frac{\xi_\gb}{\xi_\ga},
1466: \]
1467: and since the right hand side does not depend on $\gga$, then the
1468: limit exists as $N\to\infty$. It follows that for $\ell\in\ga$, $k+\ell\in\gb$:
1469: \[
1470: \lim_{N\to \infty}
1471: \bP^\rc_{N-\ell,\theta_\ell\go} \left( \tau_1 =k \right)\, = \, M_{\ga,\gb}(k)
1472: \, e^{-\tf_\go k} \, \frac{\xi_{\gb}}{\xi_\ga} \, = \,
1473: \Gamma^>_{\ga,\gb}(k).
1474: \]
1475: By the Markov property of $\bP^\rc_{N, \go}$ this yields
1476: \[
1477: \lim_{N\to \infty}
1478: \bP^\rc_{N, \go} \left( \tau_1 =k_1, \ldots, \tau_j=k_j \right)\, =\,
1479: \prod_{i=1}^j \Gamma^>_{[k_{i-1}],[k_i]}(k_i-k_{i-1}),
1480: \qquad k_0:=0.
1481: \]
1482: 
1483: \smallskip
1484: \noindent
1485: The argument for $\bP_{N,\go}^\rf$ goes along the very same line:
1486: by (\ref{eq:relfc}),
1487: %
1488: \begin{align*}
1489: & e^{-\tf_\go N} \, Z_{N-k,\theta_k\go}^\rf =
1490: e^{-\tf_\go N} \sum_{t=0}^{N-k} \cZ_{ [k],[N-t]}(N-k-t) \: P(t) \,
1491: \exp\left(\tPhi_{[N-t],[N]}(t) \right) \\
1492: & \qquad = e^{-\tf_\go k}\; \sum_{\eta \in \bbS}
1493: \sum_{t=0}^{N-k} e^{-\tf_\go \, t} \, P(t)
1494: \left[\exp\left(\tPhi_{\eta,[N]}(t) \right)
1495: \, e^{-\tf_\go \, (N-k-t)} \cZ_{ [k],\eta}(N-k-t)\right].
1496: \end{align*}
1497: %
1498: Since by~\eqref{eq:as_Z_loc} the expression in brackets converges
1499: as $N\rightarrow \infty$ and $N\in[t]+\eta$, we obtain
1500: \begin{align*}
1501: \exists \limtwo{N\to\infty}{N\in\gamma}
1502: e^{-\tf_\go N} \, Z_{N-k,\theta_k\go}^\rf \, = \,
1503: \xi_{[k]} \, e^{-\tf_\go k}
1504: \Bigg( \frac{T}{\mu} \, \sum_{\eta \in \bbS}
1505: \sumtwo{t \in \N}{[t]=\gamma - \eta} e^{-\tf_\go \, t} \, P(t) \,
1506: \exp\left(\tPhi_{\eta,\, \gamma}(t) \right)
1507: \zeta_\eta \Bigg) \,.
1508: \end{align*}
1509: Observe that the term in parenthesis is just a function of~$\gamma$. Having found
1510: the precise asymptotics of $Z_{N,\go}^\rf$, we can argue as for $\bP_{N,\go}^\rc$
1511: to conclude the proof.
1512: \qed
1513: 
1514: 
1515: 
1516: \smallskip
1517: \subsection{The critical regime $(\delta^\go=1)$}
1518: 
1519: 
1520: We prove point (3) of Theorem \ref{th:infvol}. More precisely, we
1521: prove the following:
1522: %
1523: \begin{proposition}\label{pr:infvol3}
1524: If $\delta^\go=1$ then
1525: the polymer measures $\bP^\rf_{N, \go}$ and $\bP^\rc_{N, \go}$
1526: converge as $N\to\infty$ to the same limit
1527: $\bP_\go$, under which $([\tau_k],\tau_k-\tau_{k-1})_{k\in\bbN}$
1528: is a Markov renewal process with semi-Markov kernel
1529: $(\Gamma^=_{\ga,\gb}(x): \ga,\gb\in\bbS, x\in\N)$.
1530: \end{proposition}
1531: %
1532: \noindent
1533: For the definition of $\Gamma^=$ see (\ref{eq:def_Gamma=}).
1534: 
1535: \smallskip
1536: \subsubsection{Proof of Proposition \ref{pr:infvol3}.}
1537: We prove first the case $a=\rc$.
1538: By (\ref{eq:as_L_div}) and
1539: and by the asymptotics of $\cZ$ in (\ref{eq:as_Z_cri}) above, we obtain
1540: for all $k\in\ga$:
1541: \[
1542: \exists \ \limtwo{N\to\infty}{N\in\gb} \ N^{1/2}
1543: \ \cZ_{\ga,\gb}(N-k) \, = \,
1544: \frac{T^2}{2\pi} \, \frac{\xi_\ga\, \zeta_\gb}
1545: {\sum_{\gamma,\gga'} \zeta_\gamma \,
1546: L_{\gga,\gamma'}\, \xi_{\gamma'} }.
1547: \]
1548: It follows for all $\ga,\gb,\gga\in\bbS$ and $\ell\in\ga$, $m\in\gb$
1549: \[
1550: \exists \ \limtwo{N\to\infty}{N\in\gga} \
1551: \frac{Z^\rc_{N-m,\theta_m\go}}{Z^\rc_{N-\ell,\theta_\ell\go}}
1552: \, = \, \limtwo{N\to\infty}{N\in\gga} \
1553: \frac{\cZ_{\gb,\gga}(N-m)}{\cZ_{\ga,\gga}(N)}
1554: \, = \, \frac{\xi_\gb}{\xi_\ga},
1555: \]
1556: and since the right hand side does not depend on $\gga$, then the
1557: limit exists as $N\to\infty$. It follows for $\ell\in\ga$, $k+\ell\in\gb$:
1558: \[
1559: \lim_{N\to \infty}
1560: \bP^\rc_{N-\ell,\theta_\ell\go} \left( \tau_1 =k \right)\, = \, M_{\ga,\gb}(k)
1561: \, \frac{\xi_{\gb}}{\xi_\ga} \, = \,
1562: \Gamma^=_{\ga,\gb}(k).
1563: \]
1564: By the Markov property of $\bP^\rc_{N, \go}$ this yields
1565: \[
1566: \lim_{N\to \infty}
1567: \bP^\rc_{N, \go} \left( \tau_1 =k_1, \ldots, \tau_j=k_j \right)\, =\,
1568: \prod_{i=1}^j \Gamma^=_{[k_{i-1}],[k_i]}(k_i-k_{i-1}),
1569: \qquad k_0:=0.
1570: \]
1571: 
1572: \smallskip
1573: 
1574: For $\bP_{N,\go}^\rf$,
1575: by (\ref{eq:relfc}) we have for $N\in\gb$ and $k\leq N$:
1576: \[
1577: Z_{N-k,\theta_k\go}^\rf \, = \, \sum_\gga
1578: \sum_{t=0}^{N-k} \cZ_{ [k],\gga}(t) \, P(N-k-t) \,
1579: \exp\Big(\tPhi_{\gga, \, \gb}(N-k-t) \Big).
1580: \]
1581: By the previous results and using~\eqref{eq:as_tPhi} we obtain that for every~$k\in\N$
1582: %
1583: \begin{align}\label{asZf}
1584: \begin{split}
1585: \exists \; \lim_{N\to\infty,\; N\in\gb} \, Z_{N-k,\theta_k\go}^\rf \ & = \ \xi_{[k]}\, \frac{T}{2\pi} \,
1586: \frac{\sum_\eta \zeta_\eta \, \tL_{\eta,\gb}}{\sum_{\eta,\eta'} \zeta_\eta \,
1587: L_{\eta,\eta'}\, \xi_{\eta'}}
1588: \int_0^1 \frac{dt}{t^{\frac12}(1-t)^{\frac 12}}\\
1589: & = \ \xi_{[k]}\, \Bigg( \frac{T}{2} \, \frac{\sum_\eta \zeta_\eta \, \tL_{\eta,\gb}}{\sum_{\eta,\eta'} \zeta_\eta \,
1590: L_{\eta,\eta'}\, \xi_{\eta'}} \Bigg) \, .
1591: \end{split}
1592: \end{align}
1593: %
1594: To conclude it suffices to argue as in the constrained case. \qed
1595: 
1596: \smallskip
1597: \subsection{The strictly delocalized regime $(\delta^\go<1)$}
1598: \label{sec:tl_del}
1599: 
1600: We prove point (2) and the last assertion of Theorem \ref{th:infvol}.
1601: In this case the result is different
1602: according to whether $\go\in\cP^<$ or $\go\notin\cP^<$ (recall the definition
1603: \eqref{sit1}). To be more precise, there is
1604: first a weak formulation for {\sl all} $\omega$ %such that $(\delta^\go<1)$,
1605: which gives a thermodynamic
1606: limit of $\bP^a_{N,\go}$ depending on the sequence
1607: $\{N:[N]=\eta\}$ and on $a=\rf,\rc$; secondly, there
1608: is a stronger formulation only for $\go\notin\cP^<$,
1609: which says that such limits coincide for all $\eta\in\bbS$
1610: and $a=\rf,\rc$.
1611: 
1612: \smallskip
1613: 
1614: It will turn out that in the strictly delocalized regime
1615: there exists a.s. a last return to zero, i.e. the process  $(\tau_k)_{k\in\bbN}$ is defective.
1616: In order to express this with the language of Markov renewal processes,
1617: we introduce the sets
1618: $\obbS:=\bbS\cup\{\infty\}$ and $\obbN:=\bbN\cup\{\infty\}$,
1619: extending the equivalence relation to $\obbN$ by $[\infty] =\infty$.
1620: Finally we set for all $\ga,\eta\in\bbS$:
1621: \[
1622: \Lambda_{\ga,\eta}^\rc \, := \, \big[(1-B)^{-1}L\, (1-B)^{-1}\big]_{\ga,\eta}, \qquad
1623: \mu_{\ga,\eta}^\rc \, := \, \big[L\, (1-B)^{-1}\big]_{\ga,\eta},
1624: \]
1625: \[
1626: \Lambda_{\ga,\eta}^\rf \, := \, \big[(1-B)^{-1}\tL\big]_{\ga,\eta}, \qquad
1627: \mu_{\ga,\eta}^\rf \, := \, \tL_{\ga,\eta},
1628: \]
1629: and for all $\eta\in\bbS$ and $a=\rf,\rc$
1630: we introduce the semi-Markov kernel on $\obbS\times\obbN$:
1631: %
1632: \[
1633:     \Gamma_{\ga,\gb}^{\eta,a}(x) \, := \,
1634: \begin{cases}
1635:     {\displaystyle M_{\alpha,\gb}(k) \, \Lambda_{\gb,\eta}^a/\Lambda_{\ga,\eta}^a}
1636: \quad & \ga\in\bbS, \ x\in\bbN, \ \beta=[x]\in\bbS\\
1637:     \rule{0pt}{15pt}{\displaystyle \mu_{\ga,\eta}^a/\Lambda_{\ga,\eta}^a} \quad
1638: & \ga\in\bbS, \ x=\infty, \ \beta=[\infty] \\
1639:     \rule{0pt}{15pt}1 & \ga=\gb=[\infty], \ x=0 \\
1640:     \rule{0pt}{15pt}0 & {\rm otherwise}.
1641: \end{cases}
1642: \]
1643: Notice that $\Gamma^{\eta,a}$ is really a semi-Markov kernel, since for $\ga\in\bbS$:
1644: %
1645: \begin{eqnarray*}
1646: \sum_{\gb\in\obbS} \sum_{x\in\obbN} \Gamma^{\eta,a}_{\ga,\gb}(x)
1647: & = & \frac{\mu_{\ga,\eta}^a}{\Lambda_{\ga,\eta}^a} +
1648: \sum_{\gb\in\bbS} \sum_{x\in\bbN}
1649: \frac{M_{\alpha,\gb}(x) \, \Lambda_{\gb,\eta}^a}{\Lambda_{\ga,\eta}^a}
1650: = \frac{\mu_{\ga,\eta}^a}{\Lambda_{\ga,\eta}^a} + \frac 1{\Lambda_{\ga,\eta}^a}
1651: [B\cdot \Lambda^a]_{\ga,\eta}
1652: \\ \\ & = & \frac{\mu_{\ga,\eta}^a}{\Lambda_{\ga,\eta}^a} +
1653: \frac 1{\Lambda_{\ga,\eta}^a} (\Lambda_{\ga,\eta}^a-{\mu_{\ga,\eta}^a}) \, = \, 1.
1654: \end{eqnarray*}
1655: %
1656: 
1657: \smallskip
1658: We are going to prove the following:
1659: 
1660: \smallskip
1661: %
1662: \begin{proposition}\label{pr:infvol2}
1663: Let $\delta^\go<1$. Then:
1664: %
1665: \begin{enumerate}
1666: %
1667: \item for $a=\rf,\rc$,
1668: $\bP^a_{N, \go}$
1669: converges as $N\to\infty$, $[N]=\eta$ to a measure
1670: $\bP_{\go}^{a,\eta}$, under which $([\tau_k],\tau_k-\tau_{k-1})_{k\in\bbN}$
1671: is a Markov renewal process with semi-Markov kernel
1672: $(\Gamma^{\eta,a}_{\ga,\gb}(x): \ga,\gb\in\obbS, x\in\obbN)$.
1673: %
1674: \smallskip
1675: \item
1676: if $\go\notin\cP^<$, then $\bP_{\go}^{a,\eta}=:\bP_\go$ and $\Gamma^{\eta,a}=:\Gamma^<$
1677: depend neither on $\eta$ nor on $a$, and
1678: both $\bP^\rf_{N, \go}$ and $\bP^\rc_{N, \go}$
1679: converge as $N\to\infty$ to $\bP_\go$, under which $([\tau_k],\tau_k-\tau_{k-1})_{k\in\bbN}$
1680: is a Markov renewal process with semi-Markov kernel~$\Gamma^<$.
1681: %
1682: \end{enumerate}
1683: %
1684: \end{proposition}
1685: \smallskip
1686: 
1687: %
1688: \begin{rem}\label{sit2}{\rm
1689: Part (2) of Proposition~\ref{pr:infvol2} is an easy consequence of part~(1). In fact
1690: from equations \eqref{defL} and \eqref{eq:as_tPhi} it follows immediately that
1691: when $\go\notin\cP^<$ then both matrices $(L_{\ga,\gb})$ and $(\tL_{\ga,\gb})$ are constant in $\ga$, and
1692: therefore $\Lambda^a$ factorizes into a tensor product, i.e.
1693: \[
1694: \Lambda_{\ga,\eta}^a \, = \, \lambda_\ga^a \, \nu_\eta^a, \qquad
1695: \ga,\eta\in\bbS,
1696: \]
1697: where $(\lambda_\ga^a)_{\ga\in\bbS}$ and $(\nu_\ga^a)_{\ga\in\bbS}$
1698: are easily computed.
1699: But then it is immediate to check that the semi--Markov kernel~$\Gamma^{\eta,a}=:\Gamma^<$ depends
1700: neither on $\eta$ nor on $a$.}
1701: \end{rem}
1702: %
1703: 
1704: %
1705: \smallskip
1706: \subsubsection{Proof of Proposition \ref{pr:infvol2}}
1707: By the preceding Remark it suffices to prove part~(1). For all $k\in\ga$, by \eqref{eq:as_Z_del} we have
1708: %
1709: \begin{equation}\label{eq:as_Z_div}
1710: \exists \limtwo{N\to\infty}{[N]=\beta} \; N^{3/2}
1711: \ \cZ_{\ga,\gb}(N-k)
1712: \, = \, \big[(1-B)^{-1} L \, (1-B)^{-1} \big]_{\ga,\gb}
1713: \, = \,  \Lambda_{\ga,\gb}^\rc.
1714: \end{equation}
1715: %
1716: In particular, we have for all $\ga,\gb,\eta\in\bbS$ and $\ell\in\ga$, $m\in\gb$:
1717: \[
1718: \exists \limtwo{N\to\infty}{N\in\eta} \
1719: \frac{Z^\rc_{N-m,\theta_m\go}}{Z^\rc_{N-\ell,\theta_\ell\go}}
1720: \, = \, \limtwo{N\to\infty}{N\in\eta} \
1721: \frac{\cZ_{\gb,\eta}(N-m)}{\cZ_{\ga,\eta}(N)}
1722: \, = \, \frac{\Lambda_{\gb,\eta}^\rc}{\Lambda_{\ga,\eta}^\rc},
1723: \]
1724: Then by (\ref{eq:relpmz}) we get
1725: \[
1726: \limtwo{N\to\infty}{N\in\eta} \ \bP^\rc_{N, \go}(\tau_1=k)
1727: \, = \,
1728: \frac{M_{0,[k]}(k) \, \Lambda_{[k],\eta}^\rc}{\Lambda_{0,\eta}^\rc} =
1729: \Gamma^{\eta,\rc}_{0,[k]}(k).
1730: \]
1731: By the Markov property of $\bP^\rc_{N, \go}$ this generalizes to
1732: \[
1733: \limtwo{N\to\infty}{N\in\eta} \
1734: \bP^\rc_{N, \go} \left( \tau_1 =k_1, \ldots, \tau_j=k_j \right)\, =\,
1735: \prod_{i=1}^j \Gamma^{\eta,\rc}_{[k_{i-1}],[k_i]}(k_i-k_{i-1}),
1736: \qquad k_0:=0.
1737: \]
1738: 
1739: \smallskip
1740: We prove now the case $a=\rf$.
1741: Recalling (\ref{eq:relfc}) above, we see here that
1742: \[
1743: N^{1/2} \, Z_{N-k,\theta_k\go}^\rf \, = \,
1744: \sum_{t=0}^{N-k} \cZ_{ [k],[t+k]}(t) \, N^{1/2} \, P(N-k-t) \,
1745: \exp\Big(\tPhi_{[t+k],[N]}(N-k-t) \Big).
1746: \]
1747: Then by~\eqref{eq:as_tPhi} we obtain
1748: %
1749: %
1750: \begin{equation}\label{as_Z_f}
1751: \exists \ \limtwo{N\to\infty}{N\in\eta}
1752: N^{1/2} \, Z_{N-k,\theta_k\go}^\rf \, = \, \sum_{t=0}^\infty \cZ_{ [k],[t+k]}(t)\,
1753: \tL_{[t+k],\eta} \, = \, \big[(1-B)^{-1}\tL\big]_{[k],\eta}
1754: \, = \,  \Lambda_{[k],\eta}^\rf\,,
1755: \end{equation}
1756: %
1757: since
1758: %
1759: \begin{equation}\label{mmu_gb}
1760: \sum_{t=0}^\infty \cZ_{\ga,\gga}(t) \, = \,
1761: \sum_{t=0}^\infty \sum_{k=0}^\infty M^{*k}_{\ga,\gga}(t)
1762: \, = \, \sum_{k=0}^\infty B^{*k}_{\ga,\gga} \, = \,
1763: \left[(I-B)^{-1}\right]_{\ga,\gga}.
1764: \end{equation}
1765: %
1766: Arguing as for $\bP_{N,\go}^\rc$, we conclude the proof.
1767: \qed
1768: 
1769: %\medskip\noindent
1770: %Notice that $\bP_{\go,\eta}^a(\tau_1<\infty)=1-\mu_{0,\eta}^a/\Lambda_{0,\eta}^a<1$.
1771: %Therefore, in the thermodynamic limit the process $(\tau_k)_{k\in\bbN}$ is
1772: %indeed defective.
1773: 
1774: 
1775: 
1776: \smallskip
1777: \section{Scaling limits}
1778: \label{sec:scaling_limits}
1779: 
1780: In this section we prove that the measures ${\bf P}^{a}_{N,\go}$
1781: converge under Brownian rescaling. The results and proof follow
1782: very closely those of \cite{cf:DGZ} and we shall refer to this
1783: paper for several technical lemmas.
1784: 
1785: The first step is tightness of $(Q^a_{N,\go})_{N\in\N}$ in $C([0,1])$.
1786: %
1787: \begin{lemma}\label{lemma5}
1788: For any $\go$ and $a=\rc,\rf$ the sequence $(Q^a_{N,\go})_{N\in\N}$
1789: is tight in $C([0,1])$.
1790: \end{lemma}
1791: %
1792: \noindent
1793: For the standard proof we refer to Lemma 4 in \cite{cf:DGZ}.
1794: 
1795: 
1796: \smallskip\noindent
1797: In the rest of the section we prove Theorem \ref{th:scaling}.
1798: 
1799: 
1800: \smallskip
1801: \subsection{The localized regime $(\delta^\go>1)$}
1802: 
1803: We prove point (1) of Theorem \ref{th:scaling}.
1804: By Lemma~\ref{lemma5}
1805: it is enough to prove that ${\bf P}^{a}_{N,\go}(|X^N_t|>\gep)\to 0$
1806: for all $\gep>0$ and $t\in[0,1]$ and  one can obtain this
1807: estimate explicitly. We point out however that in this
1808: regime one can avoid using the compactness lemma
1809: and one can obtain a stronger result by elementary means:
1810: observe that for any $k, n \in \N$ such that  $n>1$ and $k+n\le N$, we have
1811: \begin{multline}
1812: \label{eq:insert}
1813: \bP^a_{N,\go} \big(
1814: S_k=S_{k+n}=0, \, S_{k+i}\neq 0 \text{ for } i=1, \ldots, n-1
1815: \big) \\
1816: \le\;\frac{\frac 12 \left( 1+\exp\left( \sum_{i=1}^n \left(
1817: \go_{k+i}^{(-1)}-\go_{k+i}^{(+1)}
1818: \right)
1819: \right)\right)}
1820: {Z_{n, \theta_k \go}^{\rc}}
1821: \, =:\, \widehat K_k  (n),
1822: \end{multline}
1823: and this holds both for $a=\rc$ and $a=\rf$.
1824: Inequality \eqref{eq:insert} is obtained by using the Markov
1825: property of $S$ both in the numerator and the denominator of the expression
1826: \eqref{eq:newP}
1827: defining $\bP^a_{N,\go} \left(\cdot \right)$ after having bounded
1828: $Z_{N, \go} ^a$ from below  by inserting the event $S_k=S_{k+n}=0$.
1829: Of course $\lim_{n \to \infty} (1/n)\log  \widehat K_k  (n)= -\tf _\go$
1830: uniformly in $k$ (notice that $\widehat K_{k+T}  (n) = \widehat K_k  (n)$).
1831: Therefore if we fix $\gep >0$ by the union bound we obtain (we recall that
1832: $\{\tau_j\}_j$ and $\iota_N$ were defined in Section~\ref{sec:infvol})
1833: \begin{equation*}
1834: \begin{split}
1835: \bP^a_{N,\go} \bigg(
1836: \max_{j=1,2, \ldots, \iota_N } &
1837: \tau_j-\tau_{j-1} > \left( 1+\gep \right)
1838: \log N/{\tf_\go}
1839: \bigg)\\
1840: &\qquad  \le \,
1841: \sum_{k \le N-(1+\gep) \log N/\tf_\go} \;
1842: \sum_{n> (1+\gep) \log N /\tf_\go} \widehat K_k (n) \\
1843: &\qquad \le \; N \sum_{n> (1+\gep) \log N /\tf_\go} \max_{k=0, \ldots , T-1}\widehat K_k (n)
1844: \;\le\; \frac{c}{N^{\gep}},
1845: \end{split}
1846: \end{equation*}
1847: for some $c>0$.
1848: 
1849: Let us start with the constrained case: notice that $\bP_{N, \go} ^\rc(\dd S)$--a.s.
1850: we have $\tau_{\iota_N}=N$ and hence $\max_{j \le  \iota_N }
1851: \tau_j-\tau_{j-1} \ge \max_{n=1, \ldots, N} \vert S_n \vert$, since $|S_{n+1}-S_{n}| \le 1$.
1852: Then we immediately obtain that for any $C>1/\tf_\go$
1853: \begin{equation}
1854: \label{eq:longexc}
1855: \lim_{N \to \infty}
1856: \, \bP_{N, \go} ^\rc
1857: \left( \max_{n=1, \ldots, N} \vert S_n \vert > C \log N\right) \, =\, 0,
1858: \end{equation}
1859: which is of course a much stronger statement than the
1860: scaling limit of point (1) of Theorem~\ref{th:scaling}.
1861: If we consider instead the measure $\bP_{N, \go} ^\rf$,
1862: the length of the last excursion has to be taken into account too:
1863: however, an argument very close to the one used in  \eqref{eq:insert}
1864: yields also that the last excursion is exponentially bounded
1865: (with the same exponent) and the proof of point (1) of Theorem~\ref{th:scaling} is complete.
1866: 
1867: 
1868: 
1869: \smallskip
1870: \subsection{The strictly delocalized regime $(\delta^\go<1)$}
1871: \label{4.2}
1872: 
1873: 
1874: We prove point (2) of Theorem \ref{th:scaling}. We set for $t\in\{1,\ldots,N\}$:
1875: \[
1876: D_t \, := \, \inf\{ k=1,\ldots,N: \, k> t, \ S_k=0 \},
1877: \qquad G_t \, := \, \sup\{ k=1,\ldots,N: \, k\leq t, \ S_k=0 \}.
1878: \]
1879: The following result shows that in the strictly delocalized regime,
1880: as $N\to\infty$, the visits to zero under $\bP^a_{N,\go}$ tend to
1881: be very few and concentrated at a finite distance from the origin
1882: if $a=\rf$ and from $0$ or $N$ if $a=\rc$.
1883: %
1884: \begin{lemma}\label{pro2}
1885: If $\delta^\go<1$ there exists a constant $C>0$ such that for all $L>0$:
1886: %
1887: \begin{equation}
1888: \label{eq:zeta_cgz}
1889: \limsup_{N\to\infty} \,
1890: \bP^\rf_{N,\go}\left(G_N \ge L\right)  \, \leq \, C \, L^{-1/2},
1891: \end{equation}
1892: %
1893: \begin{equation}
1894: \label{eq:zeta_cgz2}
1895: \limsup_{N\to\infty} \
1896: \bP^\rc_{N,\go}\left(G_{N/2} \ge L\right) \, \leq \, C \, L^{-1/2},
1897: \end{equation}
1898: %
1899: \begin{equation}
1900: \label{eq:zeta_cgz3}
1901: \limsup_{N\to\infty} \
1902: \bP^\rc_{N,\go}\left(D_{N/2} \le N-L\right) \, \leq \, C \, L^{-1/2}.
1903: \end{equation}
1904: %
1905: \end{lemma}
1906: \medskip\noindent
1907: Lemma~\ref{pro2} is a quantitative version of point (2) of Theorem~\ref{th:infvol}
1908: and it is a rather straightforward complement: the proof is sketched in
1909: \S~\ref{app:ther}, in particular \eqref{eq:last0}.
1910: 
1911: \smallskip
1912: \subsubsection{The signs}
1913: 
1914: In order to prove point (2) of Theorem \ref{th:scaling}, it is now enough
1915: to argue as in the proof of Theorem 9 in \cite{cf:DGZ}, with the difference
1916: that now the excursions are not necessarily in the upper half plane, i.e.
1917: the signs are not necessarily positive.
1918: So the proof is complete if we can show that there exists the
1919: limit (as~$N\to\infty$ along~$[N] = \eta$) of the probability that the
1920: process (away from $\{0,1\}$) lives in the upper half plane.
1921: In analogy with Section~\ref{sec:tl_del}, in the general case we
1922: have different limits depending on the sequence $[N] = \eta$
1923: and on $a=\rf,\rc$, while if $\go\notin\cP^<$ all such limits
1924: coincide.
1925: 
1926: We start with the constrained case: given Lemma~\ref{pro2},
1927: it is  sufficient to  show that
1928: %
1929: \begin{equation}
1930: \label{eq:pom}
1931: \exists   \ \limtwo{N\to\infty}{N\in\eta} \
1932: \bP_{N,\go}^\rc (S_{N/2} > 0) =:\texttt p^{<,\rc}_{\go,\eta}.
1933: \end{equation}
1934: %
1935: Formula \eqref{eq:pom} follows from the fact that
1936: %
1937: \begin{align*}
1938:     \bP_{N,\go}^\rc (S_{N/2} > 0) =
1939: \sum_{\ga,\gb}
1940: \sum_{x < N/2} \sum_{y>N/2}  \frac{\cZ_{0,\ga}(x) \, \rho^+_{\ga,\gb} (y-x)
1941: \, M_{\ga,\gb}(y-x) \, \cZ_{\gb,[N]}(N-y)}{\cZ_{0,[N]}(N)} \, ,
1942: \end{align*}
1943: %
1944: where for all $z\in\N$ and $\ga,\gb\in\bbS$:
1945: %
1946: \begin{equation}\label{defrho}
1947: \rho^+_{\ga,\gb} (z) \, := \, \frac{1}
1948: {1+\exp\left(- z \, h_\go +\Sigma_{\ga,\gb}\right)},
1949: \end{equation}
1950: %
1951: cf. \eqref{defsigma}.
1952: By Dominated Convergence and by (\ref{defL}) and \eqref{mmu_gb}:
1953: \[
1954: \begin{split}
1955: \exists \ \limtwo{N\to\infty}{N\in\eta} \ N^{3/2}
1956: \sum_{x < N/2} \sum_{y>N/2}  &  \cZ_{0,\ga}(x) \, \rho^+_{\ga,\gb} (y-x) \,
1957: M_{\ga,\gb}(y-x) \, \cZ_{\gb,\eta}(N-y)\\
1958: & \, = \,
1959: \big[(1-B)^{-1}\big]_{0,\ga} \, c_K \, \frac 12 \,
1960: \exp(\go ^{(0)}_{\gb}) \, \big[(1-B)^{-1}\big]_{\gb,\eta} \, .
1961: \end{split}
1962: \]
1963: By \eqref{eq:as_Z_del} we obtain \eqref{eq:pom} with
1964: %
1965: \begin{equation}
1966: \label{eq:p_om}
1967:  \texttt p^{<,\rc}_{\go,\eta} \;:=\;
1968: \frac{\sum_{\ga,\gb}\big[(1-B)^{-1}\big]_{0,\ga}\, c_K \, \frac 12 \, \exp(\go ^{(0)}_{\gb}) \,
1969: \big[(1-B)^{-1}\big]_{\gb,\eta}}{\big[(1-B)^{-1} L \, (1-B)^{-1} \big]_{0,\eta}} .
1970: \end{equation}
1971: %
1972: Observe that by \eqref{defL}:
1973: %
1974: \begin{itemize}
1975: %
1976: \item if $h_\go > 0$ then
1977: in \eqref{eq:p_om} the denominator is equal to the numerator, so that
1978: $\texttt p^{<,\rc}_{\go,\eta} =1$ for all~$\eta$.
1979: %
1980: \item if $h_\go = 0$ and $\Sigma\equiv 0$ then
1981: in \eqref{eq:p_om} the denominator is equal to {\it twice}
1982: the numerator, so that
1983: $\texttt p^{<,\rc}_{\go,\eta} =1/2$ for all~$\eta$.
1984: %
1985: \item in the remaining case, i.e. if $\go\in\cP^<$,
1986: in general~$\texttt p^{<,\rc}_{\go,\eta}$ depends on~$\eta$.
1987: %
1988: \end{itemize}
1989: %
1990: 
1991: 
1992: \medskip
1993: Now let us consider the free case. This time it is  sufficient to show that
1994: %
1995: \begin{equation}
1996: \label{eq:pom2}
1997: \exists  \ \limtwo{N\to\infty}{N\in\eta} \  \bP_{N,\go}^\rf (S_{N} > 0) =:
1998: \, \texttt p^{<,\rf}_{\go,\eta}.
1999: \end{equation}
2000: %
2001: Formula \eqref{eq:pom2} follows from the fact that
2002: \begin{equation*}
2003:     \bP_{N,\go}^\rf (S_{N} > 0) =
2004: \sum_{\ga}\sum_{x < N}  \frac{\cZ_{0,\ga}(x) \, \cdot \frac 1 2 P(N-k)
2005: }{Z^\rf_{N,\go}}\,,
2006: \end{equation*}
2007: and using \eqref{eq:relfc}, \eqref{mmu_gb} and~\eqref{eq:as_tPhi} we obtain that \eqref{eq:pom2} holds with
2008: \begin{equation}
2009: \label{eq:p_om2}
2010:  \texttt p^{<,\rf}_{\go,\eta} \, = \,
2011: \frac{\sum_{\ga}\big[(1-B)^{-1}\big]_{0,\ga}\, c_K}
2012: {\big[(1-B)^{-1} \tL\big]_{0,\eta}} .
2013: \end{equation}
2014: %
2015: Again, observe that by \eqref{eq:as_tPhi}:
2016: %
2017: \begin{itemize}
2018: %
2019: \item if $h_\go > 0$ then
2020: in \eqref{eq:p_om2} the denominator is equal to the numerator and
2021: $\texttt p^{<,\rf}_{\go,\eta} =1$ for all~$\eta$.
2022: %
2023: \item if $h_\go = 0$ and $\Sigma\equiv 0$ then
2024: in \eqref{eq:p_om2} the denominator is equal to {\it twice}
2025: the numerator, so that
2026: $\texttt p^{<,\rf}_{\go,\eta} =1/2$ for all~$\eta$.
2027: %
2028: \item in the remaining case, i.e. if $\go\in\cP^<$,
2029: in general~$\texttt p^{<,\rf}_{\go,\eta}$ depends on~$\eta$
2030: and is different from $\texttt p^{<,\rc}_{\go,\eta}$.
2031: %
2032: \end{itemize}
2033: %
2034: 
2035: 
2036: 
2037: 
2038: 
2039: \smallskip
2040: \subsection{The critical regime $(\delta^\go=1)$}
2041: 
2042: In this section we prove point~(3) of Theorem~\ref{th:scaling}.
2043: As in the previous section, we first determine the the asymptotic behavior of the
2044: zero level set of the copolymer and then we pass to the study of the signs of the excursions.
2045: 
2046: \medskip
2047: 
2048: We introduce the random closed subset $\cA^a_N$ of $[0,1]$, describing the zero
2049: set of the polymer of size~$N$ rescaled by a factor~$1/N$:
2050: \[
2051: \bbP(\cA^a_N = A/N) \, = \, p^a_{N,\go}(A), \qquad
2052: A\subseteq\{0,\ldots,N\},
2053: \]
2054: where we recall that $p^a_{N,\go}(\cdot)$ has been defined in~\S~\ref{sec:zero_set}.
2055: Let us denote by ${\cF}$ the class of \textsl{all closed subsets} of $\R^+:=[0,+\infty)$.
2056: We are going to put on~$\cF$ a topological and measurable structure, so that we can view
2057: the law of~$\cA^a_N$
2058: as a probability measure on (a suitable $\gs$--field of)~$\cF$ and we can
2059: study the weak convergence of~$\cA^a_N$.
2060: 
2061: \smallskip
2062: 
2063: We endow $\cF$ with the topology of~Matheron, cf.~\cite{cf:Matheron}
2064: and \cite[\S~3]{cf:FFM}, which is a metrizable topology. To define it,
2065: to a closed subset $F \subseteq \R^+$ we associate the closed
2066: \textsl{nonempty} subset~$\tilde F$ of the compact interval~$[0,\pi/2]$
2067: defined by $\tilde F := \arctan\!\big( F \cup \{+\infty \} \big)$.
2068: Then the metric~$\rho(\cdot, \cdot)$ we take on~$\cF$ is
2069: \begin{equation} \label{eq:Haus}
2070: \rho(F,F') \, := \, \max\bigg\{
2071: \sup_{t\in \tilde F}\, d(t,\tilde F')\,, \ \sup_{t'\in \tilde F'} \,d(t',\tilde F)\, \bigg\} \qquad \quad F,\,F' \in \cF\,,
2072: \end{equation}
2073: where $d(s,A):=\inf\{|t-s|, t\in A\}$ is the standard distance between a point and a set.
2074: We point out that the r.h.s. of~\eqref{eq:Haus} is the so--called Hausdorff metric between
2075: the compact sets~$\tilde F,\, \tilde F'$.
2076: Thus given a sequence $\{ F_n\}_n\subset{\cF}$ and $F\in{\cF}$, we say that~$F_n \to F$ in~$\cF$
2077: if and only if $\rho(F_n , F) \to 0$. We observe that this is equivalent to requiring
2078: that for each open set $G$ and each compact $K$
2079: %
2080: \begin{equation}\label{hau1}
2081: \begin{split}
2082: & \rule{0pt}{12pt} F\cap G\ne \emptyset \quad \Longrightarrow \quad F_n\cap G\ne \emptyset
2083: \ \ {\rm eventually} \\
2084: & \rule{0pt}{12pt} F\cap K = \emptyset \quad  \Longrightarrow \quad
2085: F_n \cap K = \emptyset \ \ {\rm eventually}
2086: \end{split}
2087: \;.
2088: \end{equation}
2089: %
2090: Another necessary and sufficient condition for~$F_n \to F$ is that
2091: $d(t,F_n) \to d(t,  F)$ for every~$t\in\R^+$.
2092: 
2093: This topology makes $\cF$ a separable and compact metric space~\cite[Th.~1-2-1]{cf:Matheron},
2094: in particular a Polish space. We endow $\cF$ with the Borel $\gs$--field,
2095: and by standard theorems on weak convergence
2096: we have that also the space~$\cM_1(\cF)$ of probability measures on~$\cF$ is compact.
2097: 
2098: \smallskip
2099: 
2100: The main result of this section is to show that the law of the
2101: random set~$\cA_N^a \in \cM_1(\cF)$ converges as~$N\to\infty$ to the
2102: law of the zero set of a Brownian motion $\{B(t)\}_{t\in[0,1]}$ for~$a = \rf$
2103: or of a Brownian bridge~$\{\beta(t)\}_{t\in[0,1]}$ for~$a=\rc$.
2104: %
2105: \smallskip
2106: \begin{proposition}\label{c0bm}
2107: If~$\gd^\go = 1$ then as $N\to \infty$
2108: %
2109: \begin{equation}\label{an1}
2110: {\cA}_N^\rf \ \Longrightarrow \
2111: \left\{t\in [0,1]: B(t)=0 \right\},
2112: \end{equation}
2113: %
2114: %
2115: \begin{equation}\label{an2}
2116: {\cA}_N^\rc \  \Longrightarrow \
2117: \left\{t\in [0,1]: \beta(t)=0 \right\}.
2118: \end{equation}
2119: %
2120: \end{proposition}
2121: %
2122: \medskip
2123: 
2124: The proof of Proposition \ref{c0bm} is achieved comparing the law
2125: of ${\cA}_N^\rf$ and ${\cA}_N^\rc$ with the law of a random
2126: set ${\cR}_N$ defined as follows: recalling that
2127: $\{\tau_k\}_{k\in\N}$ denotes the sequence of return times of~$S$ to
2128: zero, we set
2129: \[
2130: \cR_N \, := \, \text{range\;}\{\tau_i/N, \ i\geq 0\}
2131: \]
2132: and we look at the law~$\cR_N$ under the critical infinite volume measure~$\bP_\go$
2133: of Proposition~\ref{pr:infvol3}. Observe that under~$\bP_\go$ the process
2134: $([\tau_k],\tau_k-\tau_{k-1})_{k\in\N}$ is a Markov renewal process, whose semi--Markov kernel
2135: is given by $\Gamma^=$. The key point of the proof is given by the following result:
2136: 
2137: 
2138: \smallskip
2139: %
2140: \begin{lemma}\label{th:JB}
2141: The law of $\{\cR_N\}_N$ under $\bP_\go$ converges weakly to the law of
2142: the random set $\{ t \geq 0: B(t) =0\}$.
2143: \end{lemma}
2144: %
2145: \smallskip
2146: 
2147: The core of the proof (see Step~1 below) uses the theory of {\sl regenerative sets}
2148: and their connection with the concept of {\sl subordinator}, see
2149: \cite{cf:FFM}. However we point out that it is also possible to give a more standard proof,
2150: using tightness and checking ``convergence of the finite dimensional distributions'':
2151: this approach is outlined in \S~\ref{app:weak_conv}.
2152: 
2153: \smallskip
2154: 
2155: \noindent
2156: {\it Proof of Lemma~\ref{th:JB}} We introduce the random set
2157: \begin{equation*}
2158:     \cR_N^{(\gb)} := \text{range}\{\tau_k/N: k\geq 0, \, [\tau_k]=\beta\} \qquad \beta\in\bbS \,.
2159: \end{equation*}
2160: Notice that $\cR_N=\cup_\beta \cR_N^{(\gb)}$.
2161: Let us also recall the definitions \eqref{defkappa} and~\eqref{defTkappa}:
2162: \[
2163: \gk_0^{(\gb)} \, := \, \inf\{k\geq 0: [\tau_k] = \beta\}, \qquad
2164: \gk_{i+1}^{(\gb)} \, := \, \inf\{k> \gk_i^{(\gb)}: [\tau_k] = \beta\},
2165: \]
2166: \[
2167: T_0^{(\gb)} \, := \, \tau_{\gk_0^{(\gb)}}, \qquad
2168: T_i^{(\gb)} \, := \, \tau_{\gk_i^{(\gb)}}-\tau_{\gk_{i-1}^{(\gb)}},
2169: \qquad i\geq 1.
2170: \]
2171: Then $(T_i^{(\gb)})_{i\geq 1}$ is under~$\bP_\go$ an IID sequence, independent of
2172: $T_0^{(\gb)}$: see the discussion before~\eqref{eq:cl_ren_pr}.
2173: We divide the rest of the proof in two steps.
2174: 
2175: \medskip\noindent{\bf Step 1.}
2176: This is the main step: we prove that the law of~$\cR_N^{(\gb)}$ under~$\bP_\go$ converges
2177: to the law of $\{ t \geq 0: B(t) =0\}$. For this we follow the proof of
2178: Lemma 5 in \cite{cf:DGZ}.
2179: 
2180: Let $\{P (t)\}_{t\ge 0}$ be a Poisson process with
2181: rate $\gamma>0$, independent of $(T_i^{(\gb)})_{i\geq 0}$. Then
2182: $\sigma_t =[T_1^{(\gb)}+\cdots+T_{P(t)}^{(\gb)}]/N$
2183: forms a non decreasing CAD process with independent
2184: stationary increments and $\sigma_0=0$: in other words
2185: $\sigma=(\sigma_t)_{t\ge 0}$ is a subordinator. Notice that
2186: \[
2187: \cR_N^{(\gb)} \, := \, T_0^{(\gb)}/N \, + \, \widehat\cR_N^{(\gb)},
2188: \qquad \widehat\cR_N^{(\gb)} \,:= \,\text{range\,} \{\sigma_t: t\ge 0\}.
2189: \]
2190: Thus ${\widehat\cR}_N^{(\gb)}$ is the (closed) range of the
2191: subordinator $\sigma$, i.e. by~\cite{cf:FFM} a regenerative
2192: set. As for any Levy process, the law of
2193: $\sigma$ is characterized by the Laplace transform of the one-time
2194: distributions:
2195: \[
2196: \mathbb E\left[ \exp\left(- \lambda\sigma_t \right)\right] \, = \,
2197: \exp\left( -t \phi_N(\lambda)\right), \qquad \lambda\ge 0, \ t\ge 0,
2198: \]
2199: for a suitable function $\phi_N:[0,\infty)\mapsto[0,\infty)$, called
2200: L\'evy exponent, which has a canonical representation, the
2201: L\'evy--Khintchin formula (see e.g. (1.15) in \cite{cf:FFM}):
2202: %
2203: \begin{eqnarray*}
2204: \phi_N (\lambda) & = & \int_{(0,\infty)} \left( 1-
2205: e^{-\lambda s}\right) \, \gamma \,
2206: {\mathbb P}(T_1^{(\gb)}/N\in \, ds)
2207: \\ \\ & = & \gamma \sum_{n=1}^\infty
2208: \left( 1-\exp(-\lambda n/N)\right) \, q^{(\gb)}(n) \;.
2209: \end{eqnarray*}
2210: %
2211: Notice now that the law of
2212: the regenerative set ${\widehat\cR}_N^{(\gb)}$ is invariant under
2213: the change of time scale $\sigma _t \longrightarrow \sigma_{ct}$,
2214: for $c>0$, and in particular independent of $\gamma>0$.
2215: Since $\phi_N \longrightarrow c\,  \phi_N$ under this change of scale, we
2216: can fix $\gamma=\gamma_N$ such that $\phi_N(1)=1$ and
2217: this will be implicitly assumed from now on.
2218: By Proposition (1.14) of \cite{cf:FFM}, the law of
2219: ${\widehat\cR}_N^{(\gb)}$ is uniquely determined by $\phi_N$.
2220: 
2221: By the asymptotics of $q^{(\gb)}$ given in (\ref{eq:cb1}),
2222: one directly obtains that
2223: $\phi_N (\lambda) \to \lambda^{1/2}=:
2224: \Phi_{BM}(\lambda)$ as $N\to\infty$.
2225: It is now a matter of
2226: applying the result in \cite[\S 3]{cf:FFM} to obtain that
2227: ${\widehat\cR}_N^{(\gb)}$ converges in law to the regenerative set
2228: corresponding to~$\Phi_{BM}$. However by direct computation
2229: one obtains that the latter is nothing but the zero level set of a Brownian motion,
2230: hence~${\widehat\cR}_N^{(\gb)} \Rightarrow \{t \in [0,1]: \ B(t) = 0\}$.
2231: From the fact that $T_0^{(\gb)}/N$ tends to~$0$ a.s., the same weak convergence for
2232: $\cR_N^{(\gb)}$ follows immediately.
2233: 
2234: \medskip
2235: \noindent{\bf Step 2.}
2236: We notice now that $\cR_N=\cup_\beta \cR_N^{(\gb)}$ is the union
2237: of non independent sets. Therefore, although we know that each $\cR_N^{(\gb)}$
2238: converges in law to $\{ t \geq 0: B(t) =0\}$,
2239: it is not trivial that $\cR_N$ converges to the same limit. We start
2240: showing that for every positive $t\geq 0$, the distance
2241: between the first point in $\cR_N^{(\ga)}$ after $t$ and the first point
2242: in $\cR_N^{(\gb)}$ after $t$ converges to zero in probability.
2243: More precisely, for any closed set $F\subset[0,\infty)$ we set:
2244: %
2245: \begin{equation}\label{d_t(F)}
2246: d_t(F)  \, := \, \inf (F\cap(t,\infty)).
2247: \end{equation}
2248: %
2249: and we claim that for all $\ga,\beta\in\bbS$ and $t\geq 0$,
2250: $|d_t(\cR_N^{(\ga)})- d_t(\cR_N^{(\gb)})| \to 0$ in probability.
2251: 
2252: Recalling \eqref{eq:cl_ren_pr} and
2253: setting $q^{(\ga;\gb)}(t)=\bP_{\theta_\ga\go}(T_0^{(\gb)} =t)$, for
2254: all $\epsilon>0$:
2255: %
2256: \begin{eqnarray*} & &
2257: \bP_\go \left(d_t(\cR_N^{(\ga)})\geq d_t(\cR_N^{(\gb)})+\epsilon\right) \\ \\ & &
2258: = \sum_k \sum_{\gga} \sum_{y=0}^{\lfloor Nt\rfloor}
2259: \bP_\go (\tau_k=y, [\tau_k]=\gga) \,
2260: \sum_{z=\lfloor Nt\rfloor -y+1}^\infty \bP_{\theta_\gga \go} (T_0^{(\gb)} =z) \
2261: \bP_{\theta_\gb \go} (T_0^{(\ga)}\geq \lfloor N\epsilon\rfloor)
2262: \\ \\ & & = \sum_{\gga} \sum_{y=0}^{\lfloor Nt\rfloor}
2263: U_{0,\gga}(y) \,
2264: \sum_{z=\lfloor Nt\rfloor-y+1}^\infty q^{(\gga;\gb)} (z) \,
2265: \sum_{w=\lfloor N\epsilon\rfloor}^\infty q^{(\gb;\ga)}(w).
2266: \end{eqnarray*}
2267: %
2268: Arguing as in the proof of \eqref{eq:cb1}, it is easy to obtain
2269: the bound: $q^{(\gb;\ga)}(w)\leq C_1 \, w^{-3/2}$, and
2270: by \eqref{eq:doney}: $U_{0,\gga}(y)\leq C_2 \, y^{-1/2}$, where $C_1,\,C_2$ are positive constants.
2271: Then asymptotically
2272: \[
2273: \bP_\go \left(d_t(\cR_N^{(\ga)})\geq d_t(\cR_N^{(\gb)})+\epsilon\right)
2274: \;\leq\; \frac{C_3}{N^{1/2}} \,
2275: \bigg( \int_0^{t/T} dy \int_{(t-y)/T}^\infty dz \int_{\epsilon/T}^\infty dw
2276: \, \frac  1{y^{1/2} \, z^{3/2}\, w^{3/2}} \bigg)
2277: \]
2278: for some positive constant~$C_3$, having used the convergence of the
2279: Riemann sums to the corresponding integral. The very same computations
2280: can be performed exchanging~$\ga$ with~$\gb$, hence the claim is proven.
2281: 
2282: Now notice that $d_t(\cR_N) = \min_{\ga \in \bbS} d_t(\cR_N^{(\ga)})$, and
2283: since $\bbS$ is a finite set we have that also
2284: $|d_t(\cR_N) - d_t(\cR_N^{(\gb)})| \to 0$ in probability for any fixed $\gb \in \bbS$.
2285: Since we already know that~$\cR_N^{(\gb)}$ converges weakly to the law
2286: of~$\{t \ge 0: B(t) = 0\}$, the analogous statement for~$\cR_N$ follows
2287: by standard arguments. More precisely,
2288: let us look at $(\cR_N, \cR_N^{(\gb)})$ as a random element of the space~$\cF \times \cF$:
2289: by the compactness of~$\cF$ it suffices to take any convergent subsequence
2290: $(\cR_{k_n}, \cR_{k_n}^{(\gb)}) \Rightarrow (\mathfrak B, \mathfrak C)$
2291: and to show that~$\bbP(\mathfrak B \neq \mathfrak C) = 0$.
2292: By the Portmanteau Theorem it is sufficient to prove that
2293: $\lim_{N\to\infty} \bP_\go (\cR_{N} \neq \cR_{N}^{(\gb)}) = 0$,
2294: and this is an immediate consequence of the decomposition
2295: \begin{equation*}
2296:     \big\{ \cR_{N} \neq \cR_{N}^{(\gb)} \big\} \;=\; \union_{t \in \Q^+} \union_{n \in \N}
2297:     \big\{ |d_t(\cR_N) - d_t(\cR_N^{(\gb)})| > 1/n \big\}\,,
2298: \end{equation*}
2299: which holds by the right--continuity of~$t\mapsto d_t$.\qed
2300: 
2301: 
2302: \bigskip
2303: 
2304: \noindent {\it Proof  of  (\ref{an2}).} First, we compute the
2305: Radon-Nykodim density of the law of $\cA^\rc_N\cap[0,1/2]$ with respect to
2306: the law of $\cR^{1/2}_N:=\cR_N\cap[0,1/2]$: for
2307: $F=\{t_1/N,\ldots,t_k/N\}\subset[0,1/2]$ with $0=:t_0<t_1<\cdots<t_k$
2308: integer numbers, the Radon--Nykodim derivative of the law of $\cA_N^\rc\cap[0,1/2]$
2309: with respect to the law of $\cR^{1/2}_N$ for $\cR^{1/2}_N=F$ is:
2310: \[
2311: f_N^\rc(g_{1/2}(F)) \, = \, f_N^\rc(t_k/N) \, = \,
2312: \frac{\sum_{n=N/2}^N  M_{[t_k],[n]}(n-t_k)
2313: \, \cZ_{[n],[N]}(N-n)} {\cZ_{0,[N]}(N) \ Q_{[t_k]}(N/2-t_k)}
2314: \ \frac{\xi_0}{\xi_{[t_k]} },
2315: \]
2316: where $Q_\ga(t) := \sum_\gb \sum_{s=t+1}^\infty \Gamma^=_{\ga,\gb}(s)$
2317: and for any closed set $F\subset[0,\infty)$ we set:
2318: %
2319: \begin{equation}\label{g_t(F)}
2320: g_t(F) \, := \, \sup(F\cap[0,t]).
2321: \end{equation}
2322: %
2323: 
2324: By \eqref{eq:as_Z_cri}, for all $\gep>0$ and uniformly in $g\in[0,1/2-\gep]$:
2325: %
2326: \begin{eqnarray*}
2327: f_N^\rc(g) & \sim & \frac {\sum_\gga L_{[Ng],\gga}
2328: \ \frac{T^2}{2\pi} \, \frac{\xi_\gga\, \zeta_{[N]}}
2329: {\sum_{\gamma,\gga'} \zeta_\gamma \,
2330: L_{\gga,\gamma'}\, \xi_{\gamma'}} \ T^{-1} \int_0^{1/2} y^{-1/2}\, (1-y-g)^{-3/2} \, dy}
2331: {\frac{T^2}{2\pi} \, \frac{\xi_0\, \zeta_{[N]}}
2332: {\sum_{\gamma,\gga'} \zeta_\gamma \, L_{\gga,\gamma'}\, \xi_{\gamma'}}
2333:  \, T^{-1} \, \sum_\gga L_{[Ng],\gga} \, \xi_\gga/\xi_{[Ng]}\ 2 \, (1/2-g)^{-1/2}}
2334: \ \frac{\xi_0}{\xi_{[Ng]} }
2335: \\ \\ & = & \frac{\sqrt{1/2}}{1-g} \, =: \, r(g).
2336: \end{eqnarray*}
2337: %
2338: If $\Psi$ is a bounded continuous functional on $\cF$
2339: such that $\Psi(F)=\Psi(F\cap[0,1/2])$ for all $F\in\cF$, then,
2340: setting $Z_B:=\left\{t\in [0,1]: B(t)=0 \right\}$ and
2341: $Z_\beta:=\left\{t\in [0,1]: \beta(t)=0 \right\}$, we get:
2342: \[
2343: {\mathbb E}[\Psi(Z_\beta)] \, = \, {\mathbb
2344: E}\left[\Psi(Z_B) \, r(g_{1/2}(Z_B))\right],
2345: \]
2346: see formula (49) in \cite{cf:DGZ}.
2347: By the asymptotics of $f_N^\rc$ we obtain that
2348: \[
2349: \bbE\left[\Psi(\cA^\rc_N)\right] \, = \,
2350: \bbE\left[\Psi(\cR^{1/2}_N) \, f_N^\rc(g_{1/2}(\cR^{1/2}_N))\right]
2351: \, \to \,
2352: \bbE\left[\Psi(Z_B) \, r(g_{1/2}(Z_B))\right] \, = \,
2353: \bbE\left[\Psi(Z_\beta)\right]
2354: \]
2355: i.e. $\cA^\rc_N\cap[0,1/2]$ converges to $Z_\beta\cap[0,1/2]$.
2356: Notice now that the distribution of
2357: the random set $\{1-t: \, t\in\cA^\rc_N\cap[1/2,1]\}$ under
2358: $\bP^\rc_{N,\go}$ is the same as the distribution of $\cA^\rc_N\cap[0,1/2]$
2359: under $\bP^\rc_{N,\overline\go}$, where $\overline\go_{[i]}:=\go_{[N-i]}$.
2360: Therefore we obtain that $\cA^\rc_N\cap[1/2,1]$
2361: converges to $Z_\beta\cap[0,1/2]$ and the proof is complete.
2362: 
2363: 
2364: \bigskip
2365: \noindent
2366: {\it Proof of (\ref{an1}).} By conditioning on the last zero, we see that
2367: if $\Psi$ is a bounded continuous functional on $\cF$ then:
2368: \[
2369: \bbE\left[\Psi(\cA^\rf_N)\right] \, = \, \sum_{t=0}^N
2370: \bbE\left[\Psi(\cA^\rc_t)\right] \, \frac{Z^\rc_{t,\go}}
2371: {Z^\rf_{N,\go}}
2372: \, P \left(N -t\right) \,
2373: \exp\left(\tPhi_{[t],[N]}(N -t)\right).
2374: \]
2375: We denote by $\beta^{t}$ a Brownian bridge over the interval
2376: $[0,t]$, i.e. a Brownian motion over $[0,t]$ conditioned to
2377: be 0 at time $t$, and we set
2378: $Z_{\beta^{t}}:=\left\{s\in [0,t]: \beta^{t}(s)=0 \right\}$.
2379: By (\ref{an2}), \eqref{eq:as_Z_cri} and (\ref{asZf}) we obtain
2380: as $N\to\infty$:
2381: %
2382: \begin{eqnarray*} & &
2383: \bbE\left[\Psi(\cA^\rf_N)\right] \, = \, \sum_{t=0}^N \sum_\gga
2384: 1_{(t\in\gga)} \,
2385: \bbE\left[\Psi(\cA^\rc_t)\right] \, \frac{Z^\rc_{t,\go}}
2386: {Z^\rf_{N,\go}}
2387: \, P \left(N -t\right) \,
2388: \exp\left(\tPhi_{\gga,[N]}(N -t)\right)
2389: \\ \\ & & \sim \int_0^1 \bbE[\Psi(Z_{\beta^{t}})] \,
2390: \frac 1{t^{\frac 12}(1-t)^{\frac 12}} \, dt \cdot \sum_\gga \frac 1T \,
2391: \frac{T^2}{2\pi} \frac{\xi_0 \, \zeta_\gga}{\sum_{\eta,\eta'} \zeta_\eta \,
2392: L_{\eta,\eta'}\, \xi_{\eta'}} \, \frac{\tL_{\gga,[N]}}{\xi_0 \, \frac T2 \,
2393: \frac{\sum_\eta \zeta_\eta \, \tL_{\eta,[N]}}{\sum_{\eta,\eta'} \zeta_\eta \,
2394: L_{\eta,\eta'}\, \xi_{\eta'}}}
2395: \\ \\ & & = \int_0^1 \bbE[\Psi(Z_{\beta^{t}})] \,
2396: \frac 1{\pi \, t^{\frac 12} (1-t)^{\frac 12} }\, dt \, = \, \bbE[\Psi(Z_B)].
2397:  \qed
2398: \end{eqnarray*}
2399: %
2400: 
2401: 
2402: \smallskip
2403: \subsubsection{The signs}
2404: 
2405: \medskip\noindent
2406: To complete the proof of point~(3) of Theorem~\ref{th:scaling} in the critical
2407: case $(\gd^\go = 1)$ we follow closely the proof given in Section~8 of~\cite{cf:DGZ}.
2408: We have already proven the convergence of the set of zeros and we have to
2409: ``paste'' the excursions. From Section~\ref{sec:infvol} we know that, {\sl conditionally
2410: on the zeros}:
2411: \begin{itemize}
2412: \item the signs $\{\gs_k\}_k$ and the absolute values $\{e_k(\cdot)\}_k$ of the excursions are independent;
2413: \item the (conditional) law of $e_k(\cdot)$ is the same as under the original random walk measure~$\bP$.
2414: \end{itemize}
2415: Furthermore, the weak convergence under diffusive rescaling on~$e_k(\cdot)$ towards
2416: the Brownian excursion~$e(\cdot)$ follows by the arguments described in~\cite{cf:DGZ}.
2417: Then it only remains to concentrate on the signs.
2418: 
2419: \smallskip
2420: 
2421: We start with the constrained case: we are going to show that for all $t\in(0,1)$
2422: %
2423: \begin{equation}
2424: \label{eq:pom3}
2425: \exists   \ \lim_{N\to\infty} \
2426: \bP_{N,\go}^\rc (S_{\lfloor tN\rfloor} > 0)
2427: \, =: \, \texttt p^=_\go,
2428: \end{equation}
2429: %
2430: and the limit is independent of $t$. We point out that actually we should fix the extremities
2431: of the excursion embracing~$t$, that is we should rather prove that
2432: \begin{equation} \label{eq:fixing}
2433:     \lim_{N\to\infty} \bP_{N,\go}^\rc \big(S_{\lfloor tN\rfloor} > 0\,,\; G_{\lfloor tN \rfloor}/N \in (a,a+\gep)\,,\; D_{\lfloor tN \rfloor}/N \in (b,b+\gep) \big) \;=\; \texttt p^=_\go\,,
2434: \end{equation}
2435: for $a < t < b$ and~$\gep>0$ (recall the definition of~$G_t$ and~$D_t$ in~\S~\ref{4.2}).
2436: However to lighten the exposition we will stick to~\eqref{eq:pom3}, since proving~\eqref{eq:fixing}
2437: requires only minor changes.
2438: 
2439: We have, recalling
2440: \eqref{defrho}:
2441: %
2442: \begin{align*}
2443:     \bP_{N,\go}^\rc (S_{\lfloor tN\rfloor} > 0) =
2444: \sum_{\ga,\gb}
2445: \sum_{x < \lfloor tN\rfloor} \sum_{y>\lfloor tN\rfloor}
2446: \frac{\cZ_{0,\ga}(x) \, \rho^+_{\ga,\gb} (y-x)
2447: \, M_{\ga,\gb}(y-x) \, \cZ_{\gb,[N]}(N-y)}{\cZ_{0,[N]}(N)} \, .
2448: \end{align*}
2449: %
2450: By Dominated Convergence and by \eqref{eq:as_Z_cri}:
2451: \[
2452: \begin{split} &
2453: \exists \ \limtwo{N\to\infty}{N\in\eta} \ N^{1/2}
2454: \sum_{x < \lfloor tN\rfloor} \sum_{y>\lfloor tN\rfloor}
2455: \cZ_{0,\ga}(x) \, \rho^+_{\ga,\gb} (y-x) \,
2456: M_{\ga,\gb}(y-x) \, \cZ_{\gb,\eta}(N-y)\\
2457: & \, = \,  \frac 1{T^2} \int_0^t dx \int_t^1 dy \, [x(y-x)^3(1-y)]^{-\frac 12}
2458: \ \left(\frac{T^2}{2\pi}\right)^2 \, \frac{\xi_0\, \zeta_\ga
2459: \, \xi_\gb \, \zeta_\eta}
2460: {(\sum_{\gamma,\gga'} \zeta_\gamma \,
2461: L_{\gga,\gamma'}\, \xi_{\gamma'})^2}
2462: \, c_K \, \frac 12 \, \exp(\go ^{(0)}_{\gb})
2463: \end{split}
2464: \]
2465: see (\ref{defL}). We obtain \eqref{eq:pom3} with
2466: %
2467: \begin{equation}
2468: \label{eq:p_om3}
2469:  \texttt p^=_\go \, \;:=\;
2470: \frac{\sum_{\ga,\gb} \zeta_\ga \, c_K \, \frac 12 \,
2471: \exp(\go ^{(0)}_{\gb}) \, \xi_{\gb}}
2472: {\sum_{\ga,\gb} \zeta_\ga \,
2473: L_{\ga,\gb}\, \xi_{\gb}}
2474: \end{equation}
2475: %
2476: Observe the following: by (\ref{defL}),
2477: %
2478: \begin{itemize}
2479: %
2480: \item if $h_\go > 0$ then
2481: in \eqref{eq:p_om3} the denominator is equal to the numerator, so that
2482: $\texttt p^=_\go =1$.
2483: %
2484: \item if $h_\go = 0$ and $\Sigma\equiv 0$ then
2485: in \eqref{eq:p_om3} the denominator is equal to {\it twice}
2486: the numerator, so that $\texttt p^=_\go =1/2$.
2487: %
2488: \end{itemize}
2489: %
2490: 
2491: 
2492: \medskip
2493: Now let us consider the free case. We are going to show that
2494: for all $t\in(0,1]$:
2495: %
2496: \begin{equation}
2497: \label{eq:pom4}
2498: \exists  \ \limtwo{N\to\infty}{[N]=\eta}
2499: \bP_{N,\go}^\rf (S_{\lfloor tN\rfloor} > 0) \, = \,
2500: \left( 1 -\frac{2\, \arcsin\sqrt t}\pi\right) \texttt p^=_\go
2501: \, + \, \frac{2\, \arcsin\sqrt t}\pi \,
2502: \texttt q^=_{\go,\eta}
2503: \, =: \, \texttt p^{=,\rf}_{\go,\eta}(t)\,,
2504: \end{equation}
2505: %
2506: where $\texttt p^=_\go$ is the same as above, see~\eqref{eq:p_om3},
2507: while $\texttt q^=_{\go,\eta}$ is defined in \eqref{eq:p_om4} below.
2508: We stress again that we should actually fix the values of~$G_{\lfloor tN \rfloor}$
2509: and~$D_{\lfloor tN \rfloor}$ like in~\eqref{eq:fixing}, proving that
2510: the limiting probability is either~$\texttt p^=_\go$ or~$\texttt q^=_{\go,\eta}$
2511: according to whether~$D_{\lfloor tN \rfloor} \le N$ or~$D_{\lfloor tN \rfloor} > N$,
2512: but this will be clear from the steps below.
2513: Formula \eqref{eq:pom4} follows from the fact that
2514: \[
2515: \begin{split}
2516:     \bP_{N,\go}^\rf (S_{\lfloor tN\rfloor} > 0) = &
2517: \sum_{\ga,\gb}\sum_{x < \lfloor tN\rfloor}
2518: \sum_{y>\lfloor tN\rfloor}  \frac{\cZ_{0,\ga}(x) \,
2519: \rho^+_{\ga,\gb} (y-x) \, M_{\ga,\gb}(y-x) \,
2520: Z^\rf_{N-y,\theta_{[y]}\go}}{Z^\rf_{N,\go}}
2521: \\ & +  \sum_{\ga}\sum_{x < \lfloor tN\rfloor} \frac{\cZ_{0,\ga}(x) \,
2522: \rho^+_{\ga,[N]} (N-x) \, P(N-x) \, \exp\left(\tPhi_{[x],[N]}(N-x)\right)}
2523: {Z^\rf_{N,\go}} \; .
2524: \end{split}
2525: \]
2526: By \eqref{asZf}, letting $N\to\infty$,
2527: the first term in the r.h.s. converges to:
2528: \[
2529: \begin{split} &
2530: \int_0^t \frac{dx}{x^{\frac 12}} \int_t^1 \frac{dy}{(y-x)^{\frac 32}} \ \cdot
2531: \\ & \ \cdot \ \sum_{\ga,\gb}
2532: \frac 1{T^2} \, \frac{T^2\, \xi_0\, \zeta_\ga}
2533: {2\pi \sum_{\gamma,\gga'} \zeta_\gamma \, L_{\gga,\gamma'}\, \xi_{\gamma'}}
2534: \ c_K \, \frac 12 \, \exp(\go ^{(0)}_{\gb})
2535: \ \frac{\xi_\gb \, \frac{T}{2} \, \sum_\gga \zeta_\gga \, \tL_{\gga,\eta}}
2536: {\sum_{\gamma,\gga'} \zeta_\gamma \, L_{\gga,\gamma'}\, \xi_{\gamma'}}
2537: \cdot \frac{\sum_{\gamma,\gga'} \zeta_\gamma \, L_{\gga,\gamma'}\, \xi_{\gamma'}}
2538: {\xi_0\, \frac{T}{2} \, \sum_\gga \zeta_\gga \, \tL_{\gga,\eta}}
2539: \\ & = \left( 1 -\frac{2\, \arcsin\sqrt t}\pi\right) \cdot \texttt p^=_\go
2540: \end{split}
2541: \]
2542: while letting $N\to\infty$ with $[N]=\eta$, the second term converges to
2543: \[
2544: \begin{split} &
2545: \int_0^t \frac{dx}{x^{\frac 12}(1-x)^{\frac 12}}  \
2546: \frac 1T \, \sum_\ga \frac{T^2\, \xi_0\, \zeta_\ga}
2547: {2\pi \sum_{\gamma,\gga'} \zeta_\gamma \, L_{\gga,\gamma'}\, \xi_{\gamma'}}
2548: \ c_K
2549: \cdot \frac{\sum_{\gamma,\gga'} \zeta_\gamma \, L_{\gga,\gamma'}\, \xi_{\gamma'}}
2550: {\xi_0 \, \frac{T}{2} \, \sum_\gga \zeta_\gga \, \tL_{\gga,\eta}}
2551: \\ & = \, \frac{2\, \arcsin\sqrt t}\pi \cdot \frac{c_K \sum_\gga \zeta_\gga}
2552: {\sum_\gga \zeta_\gga \, \tL_{\gga,\eta}} \; .
2553: \end{split}
2554: \]
2555: Therefore we obtain \eqref{eq:pom4} with:
2556: %
2557: \begin{equation}
2558: \label{eq:p_om4}
2559: \texttt q^=_{\go,\eta} \, = \,
2560: \frac{c_K \sum_\gga \zeta_\gga}
2561: {\sum_\gga \zeta_\gga \, \tL_{\gga,\eta}}  \; .
2562: \end{equation}
2563: %
2564: We observe that, by (\ref{eq:as_tPhi}):
2565: %
2566: \begin{itemize}
2567: %
2568: \item if $h_\go > 0$ or if $h_\go = 0$ and $\Sigma\equiv 0$,
2569: then $\texttt p^{=,\rf}_{\go,\eta}(t)
2570: =\texttt q^=_{\go,\eta} = \texttt p^=_\go$ for all $t$ and $\eta$
2571: %
2572: \item in the remaining case, i.e. if $\go\in\cP^=$, in general
2573: $\texttt p^{=,\rf}_{\go,\eta}(t)$ depends on $t$ and $\eta$.
2574: %
2575: \end{itemize}
2576: %
2577: 
2578: \smallskip
2579: 
2580: Now that we have proven the convergence of the probabilities of the signs
2581: of the excursion, in order to conclude the proof of point (3) of Theorem \ref{th:scaling}
2582: it is enough to argue as in the proof of Theorem 11 in \cite{cf:DGZ}.
2583: 
2584: 
2585: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2586: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2587: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2588: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2589: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2590: 
2591: \smallskip
2592: \section{Appendix}
2593: 
2594: \label{app_cgz}
2595: 
2596: \smallskip
2597: \subsection{An asymptotic result}
2598: \label{app:qb}
2599: 
2600: We are going to prove that equation \eqref{eq:cb1} holds true.
2601: Before starting, let us recall an elementary fact about Markov chains.
2602: Let $Q_{\ga,\gb}$ denote the transition matrix of an irreducible, positive
2603: recurrent Markov chain, and let us introduce the matrix $Q^{(\gamma)}$ and
2604: the (column) vector~$|\gamma\rangle$ defined by
2605: \begin{equation*}
2606:     \big[ Q^{(\gamma)} \big] _{\ga,\gb} \;:=\;
2607: Q_{\ga,\gb} \, \ind_{(\gb \neq \gamma)} \qquad \quad
2608: \big[ |\gamma\rangle \big]_\ga := \ind_{(\ga = \gamma)} \,.
2609: \end{equation*}
2610: Since for any~$\gamma$ the matrix $Q^{(\gamma)}$ has spectral radius
2611: strictly smaller than~$1$, we can define the geometric series
2612: \begin{equation*}
2613:     (1-Q^{(\gamma)})^{-1} \;:=\; \sum_{k=0}^\infty \big( Q^{(\gamma)} \big) ^k \,.
2614: \end{equation*}
2615: The interesting point is that, for every~$\gamma$, the row vector
2616: $\langle \gamma |\cdot (1-Q^{(\gamma)})^{-1}$ is (a multiple of) the left
2617: Perron--Frobenius eigenvector of the matrix~$Q$ (by $\langle \gamma |$ we
2618: denote the transposed of $| \gamma \rangle$). Similarly the column vector
2619: $(1-Q^{(\gamma)})^{-1} \cdot Q \cdot | \gamma \rangle$ is (a multiple of)
2620: the right Perron--Frobenius eigenvector of~$Q$. More precisely we have
2621: \begin{equation} \label{eq:markov_chains_fact}
2622:     \big[ \langle \gamma | \cdot (1-Q^{(\gamma)})^{-1} \big]_{\ga}=
2623:      \frac{\nu_\ga}{\nu_\gamma} \qquad \quad \big[ (1-Q^{(\gamma)})^{-1}
2624: \cdot Q \cdot | \gamma \rangle \big]_{\ga} = 1\,,
2625: \end{equation}
2626: where $\{\nu_\ga\}_\ga$ is the invariant measure of the chain, that is
2627: $\sum_\ga \nu_\ga Q_{\ga,\gb} = \nu_\gb$ and $\sum_\ga \nu_\ga = 1$.
2628: Equation \eqref{eq:markov_chains_fact} can be proved by exploiting its
2629: probabilistic interpretation in terms of expected number of visits to
2630: state~$\ga$ before the first return to site~$\gamma$, see~\cite[\S~I.3]{cf:Asm}.
2631: 
2632: \medskip
2633: 
2634: Next we turn to the asymptotic behavior of $q^{(\gb)}(x)$, giving the
2635: law of $T^{(\gb)}_0$ under~$\bbP_\gb$ (recall the notations introduced in \S~\ref{sec:as_Z_cri}).
2636: With a standard renewal
2637: argument, we can express it as
2638: \begin{equation} \label{eq:conv_qb}
2639:     q^{(\gb)}(x) \;=\; \sum_{y=0}^{x-1}
2640: \sum_{\gamma \in \bbS} \; V^{(\gb)}_{\gb,\gamma}(y) \;
2641: \Gamma^=_{\gamma,\gb}(x-y) \;=\; \big( V^{(\gb)} * \Gamma^= \big)_{\gb,\gb}(x) \,,
2642: \end{equation}
2643: where the kernel $V^{(\gb)}$ is defined by
2644: \begin{equation*}
2645:     V^{(\gb)}_{\ga,\gamma}(x) \;=\; \sum_{k=0}^\infty
2646: \big[ \big( \Gamma^{(\gb)} \big) ^{*k} \big]_{\ga,\gamma}(x)\,,
2647: \end{equation*}
2648: and we have set $\Gamma^{(\gb)}_{\ga,\gamma}(x) :=
2649: \Gamma^=_{\ga,\gamma}(x) \ind_{(\gamma \neq \gb)}$. Let us look more closely
2650: at both terms in the r.h.s. of~\eqref{eq:conv_qb}.
2651: \begin{itemize}
2652: \item For the semi--Markov kernel $\Gamma^=$, recall its
2653: definition~\eqref{eq:def_Gamma=}, the asymptotic behavior as
2654: $x\to\infty$, $[x] = \gb-\gamma$ is given by
2655: \begin{equation} \label{eq:as_Gamma=}
2656:     \Gamma^=_{\gamma,\gb}(x) \;\sim\; \frac{\tL_{\gamma,\gb}}{x^{3/2}}
2657: \qquad \quad \tL_{\gamma,\gb} := L_{\gamma,\gb} \frac{\xi_\gb}{\xi_\gamma}\,.
2658: \end{equation}
2659: Moreover, we have that
2660: \begin{equation} \label{eq:sum_Gamma=}
2661:     \sum_{x\in\N} \Gamma^=_{\gamma,\gb}(x) \;=\; B_{\gamma,\gb}
2662: \frac{\xi_\gb}{\xi_\gamma} \;=:\; \tB_{\gamma,\gb} \,.
2663: \end{equation}
2664: \item On the other hand, for the kernel $V^{(\gb)}$ we can apply the
2665: theory developed in~\S~\ref{sec:as_Z_del} for the case~$\gd^\go < 1$, because the matrix
2666: \begin{equation*}
2667:     \sum_{x\in\N} \Gamma^{(\gb)}_{\ga,\gamma}(x) \;=\; \big[ \tB^{(\gb)} \big]_{\ga,\gamma}
2668: \end{equation*}
2669: has Perron--Frobenius eigenvalue strictly smaller than~$1$ (we recall the
2670: convention $[ Q^{(\gb)} ]_{\ga,\gamma} := Q_{\ga,\gamma} \ind_{(\gamma \neq \gb)}$
2671: for any matrix~$Q$). Since
2672: \begin{equation*}
2673:     \Gamma^{(\gb)}_{\ga,\gamma}(x) \;\sim\;
2674: \frac{\big[ \tL^{(\gb)} \big]_{\ga,\gamma}}{x^{3/2}} \qquad \quad x\to\infty\,,
2675: \quad [x] = \gamma-\ga\,,
2676: \end{equation*}
2677: we can apply \eqref{eq:as_Z_del} to get that as $x\to\infty$, $[x] = \ga-\gamma$
2678: \begin{equation} \label{eq:as_V}
2679:     V^{(\gb)}_{\ga,\gamma}(x) \;\sim\;
2680: \Big( \big[(1-\tB^{(\gb)})^{-1} \tL^{(\gb)} (1-\tB^{(\gb)})^{-1} \big]_{\ga,\gamma} \Big)
2681: \,\frac{1}{x^{3/2}}\,.
2682: \end{equation}
2683: Moreover applying an analog of \eqref{eq:sum} we get that
2684: \begin{equation} \label{eq:sum_V}
2685:     \sum_{y \in \N} V^{(\gb)}_{\ga,\gamma}(y)=
2686:     \sum_{k=0}^\infty \big[ \big( \tB^{(\gb)} \big)^k\big]_{\ga,\gamma}
2687:     =
2688:     \big[ (1-\tB^{(\gb)})^{-1} \big]_{\ga,\gamma} \,.
2689: \end{equation}
2690: \end{itemize}
2691: 
2692: \smallskip
2693: 
2694: We are finally ready to get the asymptotic behavior of~$q^{(\gb)}$.
2695: As both $V^{(\gb)}$ and $\Gamma^=$ have a $x^{-3/2}$--like tail, it is easy
2696: to check from~\eqref{eq:conv_qb} that as $x\to\infty$, $x \in T\N$
2697: \begin{equation*}
2698:     q^{(\gb)}(x) \;\sim\;
2699: \sum_{\gamma \in \bbS} \Bigg\{ \bigg( \sum_{y\in\N} V^{(\gb)}_{\gb,\gamma}(y) \bigg)
2700: \Gamma^=_{\gamma,\gb} (x)  \;+\;  V^{(\gb)}_{\gb,\gamma}(x) \bigg( \sum_{y\in\N}
2701: \Gamma^=_{\gamma,\gb} (y) \bigg) \Bigg\}\,,
2702: \end{equation*}
2703: and applying \eqref{eq:sum_V}, \eqref{eq:as_Gamma=}, \eqref{eq:as_V} and \eqref{eq:sum_Gamma=}
2704: we get that $q^{(\gb)}(x) \sim c_\gb / x^{3/2}$ as $x\to\infty$, $x \in T\N$, with
2705: \begin{align*}
2706:     c_\gb &\;=\; \Big[ (1-\tB^{(\gb)})^{-1} \cdot \tL \Big]_{\gb,\gb} \;+\;
2707: \Big[ (1-\tB^{(\gb)})^{-1} \cdot \tL^{(\gb)} \cdot (1-\tB^{(\gb)})^{-1} \cdot \tB \Big]_{\gb,\gb} \\
2708:     & \;=\; \Big[ (1-\tB^{(\gb)})^{-1} \cdot \tL \cdot (1-\tB^{(\gb)})^{-1} \cdot \tB \Big]_{\gb,\gb}\\
2709:     &\;=\; \langle \gb | \cdot  (1-\tB^{(\gb)})^{-1} \cdot \tL \cdot
2710: (1-\tB^{(\gb)})^{-1} \cdot \tB   \cdot | \gb \rangle\,.
2711: \end{align*}
2712: To obtain the second equality we have used the fact that
2713: \begin{equation*}
2714:     \Big[ (1-\tB^{(\gb)})^{-1} \cdot \tB \Big]_{\gb,\gb} \;=\;
2715: \Big[ \langle \gb | \cdot (1-\tB^{(\gb)})^{-1} \cdot \tB \Big]_\gb \;=\; 1\,,
2716: \end{equation*}
2717: which follows from \eqref{eq:markov_chains_fact} applied to the matrix~$Q=\tB$.
2718: Again from~\eqref{eq:markov_chains_fact} we get
2719: \begin{equation*}
2720:     c_\gb \;=\; \frac{1}{\nu_\gb} \sum_{\ga,\gamma \in \bbS} \nu_\ga \tL_{\ga,\gamma}\,,
2721: \end{equation*}
2722: where $\{\nu_\ga\}_\ga$ is the invariant measure (that is the
2723: normalized left Perron--Frobenius eigenvector) of the matrix~$\tB$.
2724: However from the definition~\eqref{eq:sum_Gamma=} of~$\tB$ it is
2725: immediate to see that $\{\nu_\ga\} = \{\zeta_\ga\, \xi_\ga\}$, and
2726: recalling the definition~\eqref{eq:as_Gamma=} of~$\tL$ we obtain the
2727: expression for~$c_\gb$ we were looking for:
2728: \begin{equation}\label{eq:cb}
2729:      c_\gb \;=\; \frac 1{\zeta_\gb\, \xi_\gb}
2730: \sum_{\ga,\gamma} \zeta_\ga \, L_{\ga,\gamma}\, \xi_\gamma.
2731: \end{equation}
2732: 
2733: 
2734: 
2735: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2736: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2738: 
2739: 
2740: \smallskip
2741: \subsection{Some computations on the thermodynamic limit measure}
2742: 
2743: \label{app:ther}
2744: 
2745: 
2746: We want now to give a description of the typical
2747: paths under $\bP^{a,\eta}_\go$ in the delocalization regime, i.e.
2748: when $\delta^\go<1$. We are going
2749: to compute the distribution of two interesting
2750: random variables under $\bP^{a,\eta}_\go$ in this case: the last return
2751: to zero and the total number of returns to zero. Other
2752: analogous computations are possible using the same procedure.
2753: 
2754: \smallskip
2755: \subsubsection{The last return to zero}
2756: We want to study the law under $\bP^{a,\eta}_\go$ of the last zero
2757: $\ell:=\sup\{i\in\N: S_i=0\}$ in the strictly delocalized regime.
2758: For simplicity we consider the case $a=\rc$, the case
2759: $a=\rf$ being completely analogous.
2760: We compute first the law of $\ell_k:=\sup\{i\leq k: S_i=0\}$ with
2761: $k\in\N$: for $x\leq k< N$ and $N\in\eta$:
2762: %
2763: \begin{equation}\label{bbb}
2764: \bP^\rc_{N, \go}(\ell_k \geq x) \, = \,
2765: \sum_{y=x}^k {\cZ_{0,[y]}(y)} \sum_{z=k+1}^N
2766: \frac{ M_{[y],[z]}(z-y) \, {\mathcal Z_{[z],\eta}(N-z)}}{\cZ_{0,\eta}(N)}
2767: \end{equation}
2768: %
2769: By (\ref{eq:as_L_div}) and (\ref{eq:as_Z_div}) we obtain:
2770: \[
2771: \limtwo{N\to\infty}{N\in\eta} \, \bP^\rc_{N, \go}(\ell_k \geq x)
2772: =
2773: \sum_{y=x}^k  \cZ_{0,[y]}(y) \left[ \sum_{z=0}^\infty
2774: \frac{ L_{[y],\eta-[z]}}{ \Lambda^\rc_{0,\eta} } \,
2775: \cZ_{\eta-[z],\eta}(z) + \sum_{z=k+1}^\infty M_{[y],[z]}(z-y) \,
2776: \frac{\Lambda^\rc_{[z],\eta}}{\Lambda^\rc_{0,\eta}}\right]
2777: \]
2778: Notice now that, by \eqref{mmu_gb}:
2779: %
2780: \begin{equation}\label{mu_gb}
2781: \sum_{z=0}^\infty L_{[y],\eta-[z]} \, \cZ_{\eta-[z],\eta}(z)
2782: \, = \, \sum_\gga L_{[y],\gga} \sum_{z=0}^\infty \cZ_{\gga,\eta}(z) \, = \,
2783: \left[ L \cdot (I-B)^{-1}\right]_{[y],\eta} \, = \, \mu^\rc_{[y],\eta}.
2784: \end{equation}
2785: %
2786: Therefore, we have proven that:
2787: \[
2788: \bP^{\rc,\eta}_\go(\ell_k\geq x) \, = \, \limtwo{N\to\infty}{N\in\eta}
2789: \ \bP^\rc_{N, \go}(\ell_k \geq x) \, =
2790: \sum_{y=x}^k \cZ_{0,[y]}(y) \left[ \frac{\mu^\rc_{[y],\eta}}{\Lambda^\rc_{0,\eta}} \,
2791: + \sum_{z=k+1}^\infty M_{[y],[z]}(z-y) \, \frac{\Lambda^\rc_{[z],\eta}}{\Lambda^\rc_{0,\eta}}\right]
2792: \]
2793: and letting $k\to\infty$ we obtain:
2794: \[
2795: \bP^{\rc,\eta}_\go(\ell\geq x) \, = \, \sum_{y=x}^\infty \cZ_{0,[y]}(y) \,
2796: \frac{\mu^\rc_{[y],\eta}}{\Lambda^\rc_{0,\eta}}.
2797: \]
2798: For the proof of Lemma \ref{pro2} above, notice for instance that by (\ref{bbb}):
2799: %
2800: \begin{eqnarray}\label{eq:last0}
2801: & &
2802: \bP^\rc_{N, \go}(G_{N/2} \geq L) \, = \,
2803: \bP^\rc_{N, \go}(\ell_{N/2} \geq L)
2804: \\ \nonumber \\ \nonumber
2805: & & \leq \,
2806: C_1 \, N^{3/2} \sum_{t=L}^{\lfloor N/2\rfloor} t^{-3/2}
2807: \sum_{k= \lfloor N/2\rfloor +1}^{N+1} (k-t)^{-3/2} \,
2808: (N+2-k)^{-3/2} \, \le \, C_2 \, L^{-1/2},
2809: \end{eqnarray}
2810: %
2811: where $C_1,C_2$ are positive constants.
2812: 
2813: \smallskip
2814: 
2815: \subsubsection{The number of returns to zero}
2816: Analogously, we want to study the law of the total number of returns to
2817: zero $\cN:=\#\{i\in\N: S_i=0\}$ under $\bP^{\rc,\eta}_{\go}$. We study first
2818: Let $\cN_K:=\#\{i: 1\leq i\leq K: S_i=0\}$ for $k\in\N$.
2819: For $k\leq K$ and $N\in\eta$:
2820: \[
2821: \bP^\rc_{N, \go}(\cN_K = k) \, = \,
2822: \sum_{x=1}^K M^{*k}_{0,[x]}(x) \sum_{y=K+1}^N
2823: \frac{M_{[x],[y]}(y-x) \, {\mathcal Z_{[y],\eta}(N-y)}}{\cZ_{0,\eta}(N)}
2824: \]
2825: Then by (\ref{eq:as_L_div}) and (\ref{eq:as_Z_div}):
2826: %
2827: \begin{eqnarray*}& &
2828: \limtwo{N\to\infty}{N\in\eta} \ \bP^\rc_{N, \go}(\cN_K = k)
2829: \\ \\
2830: & & = \sum_{x=0}^K M^{*k}_{0,[x]}(x) \left[\sum_{y=0}^\infty
2831: \frac{ L_{[x],\eta-[y]}}{\Lambda^\rc_{0,\eta}} \cZ_{\eta-[y],\eta}(y)
2832: + \sum_{z=K+1}^\infty M_{[x],[y]}(y-x)
2833: \frac{\Lambda^\rc_{[y],\eta}}{\Lambda^\rc_{0,\eta}}\right].
2834: \end{eqnarray*}
2835: %
2836: By (\ref{mu_gb}), letting $K\to\infty$ we obtain:
2837: \[
2838: \bP^{\rc,\eta}_\go(\cN = k) \, = \, \frac 1{\Lambda^\rc_{0,\eta}} \left[B^k
2839: \cdot \mu^\rc\right]_{0,\eta}.
2840: \]
2841: 
2842: 
2843: 
2844: 
2845: 
2846: 
2847: 
2848: 
2849: 
2850: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2851: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2852: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2853: 
2854: 
2855: 
2856: \smallskip
2857: 
2858: 
2859: \subsection{On the weak convergence of the critical zero set}
2860: \label{app:weak_conv}
2861: 
2862: We are going to outline an alternative proof of Lemma~\ref{th:JB},
2863: that is we are going to show that when~$\gd^\go = 1$ as~$N\to\infty$
2864: \begin{equation} \label{eq:app_goal}
2865:     \cR_{N} \text{ under } \bP_\go \;\Longrightarrow\; \{ t \geq 0: B(t) =0\}.
2866: \end{equation}
2867: To keep the notation transparent, it is convenient to denote by~$\cG_N \in \cM_1(\cF)$
2868: the image law of~$\cR_N$ under~$\bP_\go$. That is $\cG_N$~is a probability law on~$\cF$
2869: (the class of all closed subsets of~$\R^+$) defined for a measurable subset~$A \subseteq \cF$ by
2870: \begin{equation*}
2871:     \cG_N ( A) \;:=\; \bP_\go \big( \cR_N \in A \big) \,.
2872: \end{equation*}
2873: In the same way the law of~$\{ t \geq 0: B(t) =0\}$ will be denoted by~$\cG^{(BM)}$.
2874: Then we can reexpress our goal~\eqref{eq:app_goal} as
2875: \begin{equation} \label{eq:app_goal1}
2876:     \cG_N \;\Longrightarrow \; \cG^{(BM)} \;.
2877: \end{equation}
2878: 
2879: \smallskip
2880: 
2881: Remember the definition \eqref{d_t(F)} of the mapping $d_t : \cF \mapsto \R^+ \cup \{+\infty\}$.
2882: We claim that to prove \eqref{eq:app_goal1} it suffices to show that, for every~$n\in\N$ and for all
2883: $t_1, \ldots, t_n \in \R$, the law of the vector~$(d_{t_1}, \ldots, d_{t_n})$ under~$\cG_N$
2884: converges to the law of the same vector under~$\cG^{(BM)}$:
2885: \begin{equation}\label{eq:finite_dim}
2886:     (d_{t_1}, \ldots, d_{t_n}) \circ \big( \cG_N \big) ^{-1} \; \Longrightarrow \;
2887: (d_{t_1}, \ldots, d_{t_n}) \circ \big( \cG^{(BM)} \big) ^{-1} \,.
2888: \end{equation}
2889: 
2890: \smallskip
2891: 
2892: The intuitive explanation of why~\eqref{eq:finite_dim} should imply~\eqref{eq:app_goal1}
2893: is that an element $\xi \in \cF$ can be {\sl identified} with the process~$\{d_t(\xi)\}_{t\in\R^+}$,
2894: since $\xi = \{t\in\R^+:\; d_{t-}(\xi) = t\}$. Hence the convergence in~$\cM_1(\cF)$ can be read in
2895: terms of the random process~$\{d_t(\cdot)\}_{t\in\R^+}$, and using the compactness of~$\cM_1(\cF)$
2896: it turns out that~\eqref{eq:finite_dim} is indeed sufficient to ensure~\eqref{eq:app_goal1}. Let
2897: us sketch more in detail these arguments.
2898: \begin{enumerate}
2899: \item \label{point:sigma}The Borel $\gs$--field of~$\cF$ coincides with $\gs(\{d_t\}_{t\in\R^+})$,
2900: i.e. with the $\gs$-field generated by~$\{d_t\}_{t\in\R^+}$, and also with~$\gs(\{d_t\}_{t\in I})$ where~$I$
2901: is any dense subset of~$\R^+$.
2902: \item \label{point:caglad} Suppose that we are given $\{\nu_k\},\,\nu \in \cM_1(\cF)$ such that
2903: $\nu_k \Rightarrow \nu$: this fact does not entail the convergence of all the finite
2904: dimensional marginals of~$\{d_t\}$, that is \textsl{it is not true} that the law of the vector
2905: $(d_{t_1}, \ldots, d_{t_n})$ under~$\nu_k$ converges to the law of the same vector under~$\nu$,
2906: because the mappings $d_t(\cdot)$ are not continuous on~$\cF$. Nevertheless one can show that
2907: this convergence does hold for \textsl{almost all} choices of the indexes~$t_1, \ldots, t_n$.
2908: More precisely, given any measure $\nu \in \cM_1(\cF)$ there exists a subset~$I_\nu \subseteq \R^+$
2909: with $Leb(I_\nu{}^\rc)=0$ with the following property: for any sequence $\{\nu_k\}$ with
2910: $\nu_k \Rightarrow \nu$, for any~$n\in\N$ and for all $t_1, \ldots, t_n \in I_\nu$, the law of
2911: the vector~$(d_{t_1}, \ldots, d_{t_n})$ under~$\nu_k$ converges as~$k\to\infty$ to the law of
2912: the same vector under~$\nu$. This is a well-known feature of processes whose discontinuity points
2913: form a negligible set, in particular CADLAG processes: in fact the set~$I_\nu$ can be chosen as
2914: the set of~$t\in\R^+$ such that $\nu\big\{\xi:\, d_{t-}(\xi) = d_{t}(\xi)\big\} = 1$, because
2915: $d_{t-}(\xi) = d_t(\xi)$ implies that $d_t(\cdot)$ is continuous at~$\xi$.
2916: \item Since~$\cM_1(\cF)$ is compact, to prove~\eqref{eq:app_goal1} it suffices to show that any
2917: convergent subsequence of~$\{\cG_N\}_N$ converges to~$\cG^{(BM)}$. Thus we take a convergent
2918: subsequence $\cG_{k_n} \Rightarrow \nu$ for some~$\nu\in \cM_1(\cF)$ and we want to prove
2919: that~$\nu = \cG^{(BM)}$. By point~\eqref{point:caglad} there exists a dense subset~$I_\nu \subseteq \R^+$
2920: such that for $t_1, \ldots, t_n \in I_\nu$ the law of the vector~$(d_{t_1}, \ldots, d_{t_n})$
2921: under~$\cG_{k_n}$ converges to the law of the same vector under~$\nu$, and since we are assuming
2922: that~\eqref{eq:finite_dim} holds this means that the vector~$(d_{t_1}, \ldots, d_{t_n})$ has the
2923: same law under~$\nu$ and under~$\cG^{(BM)}$. This is equivalent to say that $\nu$ and $\cG^{(BM)}$
2924: coincide on the $\gs$--field $\gs(\{d_t\}_{t\in I_\nu})$, and by point~\eqref{point:sigma} it
2925: follows that indeed~$\nu = \cG^{(BM)}$.
2926: \end{enumerate}
2927: 
2928: \smallskip
2929: 
2930: Thus it only remains to show that~\eqref{eq:finite_dim} holds, and this can be done by direct
2931: computation. For simplicity we consider only the case~$n=1$ of the one--time marginals, but
2932: everything can be extended to the case~$n>1$.
2933: 
2934: For any~$t>0$ the law of~$d_t$ under~$\cG^{(BM)}$ is given by
2935: \begin{equation*}
2936:     \cG^{(BM)} \big( d_t \in \dd y \big) \;=\; \frac{t^{1/2}}{\pi \, y(y-t)^{1/2}} \,
2937: \ind_{(y > t)} \, dy \;=:\; \rho_t(y) \, dy \,,
2938: \end{equation*}
2939: see \cite{cf:RevYor}. Hence we have to show that for every~$x\in\R^+$
2940: \[
2941: \lim_{N\to\infty} \, \bP_\go \big(d_t(\cR_N) > x\big)
2942: \, = \,\int_x^\infty \rho_t(y) \, dy \,.
2943: \]
2944: We recall that $\cR_N = \text{range}\{\tau_n/N:\; n \ge 0\}$ is the range of the
2945: process~$\{\tau_n\}_{n\in\N}$ rescaled by a factor~$1/N$, and that under~$\bP_\go$
2946: the process~$\{\tau_n\}_{n\in\N}$ is a Markov--renewal process with semi--Markov
2947: kernel~$\Gamma^=_{\ga,\gb}(x)$ defined by~\eqref{eq:def_Gamma=}. We also use the
2948: notation~$U_{\ga,\gb}(x)$ for the corresponding Markov--Green function, defined
2949: by~\eqref{eq:def_U}. Then using the Markov property we get
2950: %
2951: \begin{align*}
2952: & \bP_\go \big(d_t(\cR_N) > x\big) = \sum_{k\in\N} \bP_\go \big(\tau_k \leq Nt\,,\;
2953: \tau_{k+1}> Nx\big) \\
2954: &\qquad = \sum_{\ga,\gb \in \bbS} \sum_{y=1}^{Nt} \sum_{w=Nx}^\infty \sum_{k\in\N}
2955: \bP_\go \big( \tau_k=y\,,\; [\tau_k] = \ga \big) \, \bP_{\theta_y \go} \big( \tau_1
2956: = w-y \,,\; [\tau_1] = \gb - \ga \big) \\
2957: &\qquad = \sum_{\alpha,\beta\in{\mathbb S}}
2958: \sum_{y=1}^{Nt} U_{0,\alpha}(y) \,
2959: \sum_{w=Nx}^\infty \Gamma^=_{\alpha,\beta}(w-y)
2960: \end{align*}
2961: %
2962: The asymptotic behavior of the terms appearing in the expression can be extracted
2963: from~\eqref{eq:doney} and~\eqref{eq:asympt_cgz}: the net result is that as~$z\to\infty$
2964: \begin{align*}
2965:     &\sqrt{z} \; U_{0,\alpha}(z) \; \xrightarrow{[z]=\ga} \; \frac{T^2}{2\pi} \,
2966: \frac{\zeta_\ga \xi_\ga}{\sum_{\gamma, \gamma'} \zeta_\gamma L_{\gamma,\gamma'}
2967: \xi_{\gamma '}} \;=:\; c^U_{0,\ga} \\
2968:     &z^{3/2} \; \Gamma^=_{\ga,\gb}(z) \; \xrightarrow{[z]=\gb-\ga} \;
2969: \frac{\xi_\gb}{ \xi_\ga} \, L_{\ga,\gb} \;=:\; c^\Gamma_{\ga,\gb}\,.
2970: \end{align*}
2971: Therefore we have as~$N\to\infty$
2972: \begin{align*}
2973: & \bP_\go \big(d_t(\cR_N) > x\big) \;\sim\; \sum_{\alpha,\beta\in{\mathbb S}}
2974: c^U_{0,\ga} \, c^\Gamma_{\ga,\gb}
2975: \sum_{y=1}^{Nt}  \frac{1}{\sqrt{y}} \, \ind_{([y]=\ga)}
2976: \sum_{w=Nx}^\infty \frac{1}{(w-y)^{3/2}} \, \ind_{([w]=\gb)} \\
2977: & \qquad \sim \; \frac{1}{T^2} \Bigg( \sum_{\alpha,\beta\in{\mathbb S}}
2978: c^U_{0,\ga} \, c^\Gamma_{\ga,\gb} \Bigg)
2979: \frac{1}{N^2} \sum_{s \in (0, \frac{t}{T}) \cap \frac{\Z}{N}}  \frac{1}{\sqrt{s}}
2980: \sum_{u \in (\frac{x}{T}, \infty) \cap \frac{\Z}{N}} \frac{1}{(u-s)^{3/2}}\,,
2981: \end{align*}
2982: and from the explicit expressions for~$c^U_{0,\ga},\, c^\Gamma_{\ga,\gb}$ together with
2983: the convergence of the Riemann sums to the corresponding integral we get
2984: \begin{align*}
2985:     & \exists \lim_{N\to\infty} \bP_\go \big(d_t(\cR_N) > x\big) \;=\; \frac{1}{2\pi}
2986: \int_0^{t/T} \dd s \, \frac{1}{\sqrt{s}} \int_{x/T}^\infty \dd u \, \frac{1}{(u-s)^{3/2}} \\
2987:     &\qquad \;=\; \frac{1}{\pi} \int_0^{t/T} \dd s \, \frac{1}{\sqrt{s}} \, \frac{1}{\sqrt{x/T-s}}
2988:     \;=\; \frac{1}{\pi} \int_0^{t} \dd y \, \frac{1}{\sqrt{y}} \, \frac{1}{\sqrt{x-y}}
2989:     \;=\; \int_x^\infty \dd z \, \rho_t(z) \;,
2990: \end{align*}
2991: that is what was to be proven.
2992: 
2993: 
2994: %To complete the identification of a weak limit of $\cR^{(0)}_N$
2995: %it is enough to identify the limit of $\bP_\go (d_{t_1}(\cR^{(0)}_N)>
2996: %x_1,\ldots,d_{t_n}(\cR^{(0)}_N)> x_n)$ for $t_i\leq t_{i+1}$ and $t_i\leq x_i\leq x_{i+1}$.
2997: %More generally we see that for $x_i\geq t_i$, $i=1,\ldots,n$:
2998: %
2999: %\begin{eqnarray*} & &
3000: %\bP_\go (d_{t_i}(\cR^{(0)}_N)> x_i) = \sum_{y_i=\lfloor Nx_i\rfloor+1}^\infty \prod_{i=1}^n
3001: %\bP_\go (d_{(Nt_i-y_{i-1})^+}(\cR^{(0)}_N)=y_i-y_{i-1}) \to
3002: %\\ \\ & & \int_{x_1}^\infty \cdots \int_{x_n}^\infty \prod_{i=1}^n
3003: %1_{(y_{i-1}< y_i)}  \left(
3004: %1_{(y_{i-1}<t_i)} \, \rho_{(t_i-y_{i-1})}(y_i-y_{i-1}) \, dy_i
3005: %\, + \, 1_{(y_{i-1}\geq t_i)} \, \delta_{y_{i-1}}(dy_i).
3006: %\right)
3007: %\end{eqnarray*}
3008: %
3009: %Notice that this strange formula means that: if $d_{t_{i-1}}$
3010: %is greater than $t_i$, then $d_{t_i}=d_{t_{i-1}}$ and
3011: %so there is a Dirac delta.
3012: 
3013: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3014: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3015: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3016: 
3017: 
3018: \smallskip
3019: 
3020: \subsection{A localization argument}
3021: 
3022: \label{app:loc}
3023: 
3024: Let us give a proof that for the \textsl{copolymer near a selective interface}
3025: model, described in~\S~\ref{sec:2examples}, the charge~$\go$ never belongs
3026: to~$\cP$ (see~\eqref{sit1} for the definition of~$\cP$). More precisely, we
3027: are going to show that if~$h_\go=0$ and $\Sigma\not \equiv 0$ then~$\gd^\go > 1$,
3028: that is the periodic copolymer with zero--mean, nontrivial charges is always
3029: localized. As a matter of fact this is an immediate consequence of the estimates
3030: on the critical line obtained in~\cite{cf:BG}. However we want to give here an
3031: explicit proof, both because it is more direct and because the model studied
3032: in~\cite{cf:BG} is built over the simple random walk measure, corresponding
3033: to~$p=1/2$ with the language of~\S~\ref{sec:intro_cgz}, while we consider the
3034: case~$p < 1/2$.
3035: 
3036: \smallskip
3037: 
3038: We give some preliminary notation: given an \textsl{irreducible} $T\times T$ matrix $Q_{\ga,\gb}$ with \textsl{nonnegative entries}, its Perron--Frobenius eigenvalue (= spectral radius) will be denoted by ${\tt Z}= {\tt Z}(Q)$ and the corresponding left and right eigenvectors (with any normalization) will be denoted by $\{\zeta_\ga\}, \{\xi_\ga\}$. Being a simple root of the characteristic polynomial, ${\tt Z}(Q)$ is an \textsl{analytic} function of the entries of~$Q$, and one can check that
3039: \begin{equation} \label{eq:der}
3040:     \frac{\partial {\tt Z}}{\partial Q_{\ga,\gb}}  \;=\; \frac{\zeta_\ga\,\xi_\gb}{\big( \sum_{\gamma} \zeta_\gamma \xi_\gamma \big)}\,.
3041: \end{equation}
3042: Hence ${\tt Z}(Q)$ is a strictly increasing function of each of the entries of~$Q$. We also point out a result proved by Kingman~\cite{cf:Kin}: if the matrix is a function of a real parameter $Q=Q(t)$ such that all the entries $Q_{\ga,\gb}(t)$ are \textsl{log--convex} functions of~$t$ (that is $t \mapsto \log Q_{\ga,\gb} (t)$ is convex for all~$\ga,\gb$), then also $t \mapsto {\tt Z}(Q(t))$ is a log--convex function of~$t$.
3043: 
3044: \smallskip
3045: 
3046: Next we come to the \textsl{copolymer near a selective interface} model: with reference to the general Hamiltonian~\eqref{eq:genH}, we are assuming that $\go_n^{(0)} = \tilde \go_n^{(0)} = 0$ and~$h_\go = 0$ (where~$h_\go$ was defined in~\eqref{eq:ass_h}). In this case the integrated Hamiltonian~$\Phi_{\ga,\gb} (\ell)$, see~\eqref{eq:def_Phi}, is given by
3047: \begin{equation*}
3048:     \Phi_{\ga,\gb} (\ell) = \begin{cases}
3049: 0 & \text{if} \ \ \ell = 1 \ \ \text{or} \ \ \ell \notin \gb-\ga \\
3050: \rule{0pt}{18pt}\log \Big[ \frac 12 \Big( 1 + \exp \big( \Sigma_{\ga,\gb} \big) \Big) \Big] & \text{if} \ \ \ell > 1 \ \ \text{and} \ \ \ell \in \gb-\ga
3051: \end{cases} \;.
3052: \end{equation*}
3053: We recall that the law of the first return to zero of the original walk
3054: is denoted by $K(\cdot)$, see \eqref{eq:first_ret}, and we introduce the
3055: function~$q: \bbS \to \R^+$ defined by
3056: \begin{equation*}
3057:     q(\gamma) := \sum_{x \in \N, \ [x]=\gamma} K(x)
3058: %\qquad \quad \tilde    \Sigma_{\ga,\gb} :=\begin{cases}
3059: %\Sigma_{\ga,\gb} & \text{if } \gb-\ga \neq 1\\
3060: %0 & \text{if } \gb -\ga = 1
3061: %\end{cases}
3062: \end{equation*}
3063: (notice that $\sum_\gamma q(\gamma)=1$). Then the matrix~$B_{\ga,\gb}$ defined by~\eqref{eq:def_B} becomes
3064: \begin{equation} \label{eq:B}
3065:     B_{\ga,\gb} \;=\;
3066: \begin{cases}
3067: \frac 12 \Big( 1 + \exp \big( \Sigma_{\ga,\gb} \big) \Big) \, q(\gb-\ga)  & \ \text{if} \ \ \gb-\ga \neq [1]\\
3068: \rule{0pt}{18pt}K(1) \;+\; \frac 12 \Big( 1 + \exp \big( \Sigma_{\ga,\ga + [1]} \big) \Big) \cdot \big( q([1]) - K(1) \big)  & \ \text{if} \ \ \gb-\ga = [1]
3069: \end{cases}
3070: \end{equation}
3071: By \eqref{eq:def_delta}, to prove localization we have to show that the Perron--Frobenius eigenvalue of the matrix~$(B_{\ga,\gb})$ is strictly greater than~$1$, that is ${\tt Z}(B) > 1$.
3072: 
3073: \smallskip
3074: 
3075: Applying the elementary convexity inequality $(1+\exp(x))/2 \ge \exp(x/2)$ to \eqref{eq:B} we get
3076: \begin{equation} \label{eq:step_matrix}
3077:     B_{\ga,\gb} \;\ge\; \tilde  B_{\ga,\gb} \;:=\;
3078: \begin{cases}
3079: \exp \big( \Sigma_{\ga,\gb} / 2 \big) \, q(\gb-\ga)  & \ \text{if} \ \ \gb-\ga \neq [1]\\
3080: \rule{0pt}{15pt}K(1) \;+\; \exp \big( \Sigma_{\ga,\ga + [1]} / 2 \big) \cdot \big( q([1]) - K(1) \big)  & \ \text{if} \ \ \gb-\ga = [1]
3081: \end{cases} \;.
3082: \end{equation}
3083: By hypothesis $\Sigma_{\ga_0,\gb_0} \neq 0$ for some $\ga_0,\gb_0$, therefore the inequality above is strict for $\ga=\ga_0$, $\gb=\gb_0$. We have already observed that the P--F eigenvalue is a strictly increasing function of the entries of the matrix, hence ${\tt Z}(B) > {\tt Z}(\tilde B)$. Therefore it only remains to show that ${\tt Z}(\tilde B)\ge 1$, and the proof will be completed.
3084: 
3085: Again an elementary convexity inequality applied to the second line of~\eqref{eq:step_matrix} yields
3086: \begin{equation} \label{eq:Bhat}
3087:     \tilde B_{\ga,\gb} \;\ge\; \widehat  B_{\ga,\gb} \;:=\; \exp \big(\, c(\gb-\ga) \, \Sigma_{\ga,\gb} / 2 \, \big) \cdot q(\gb-\ga)
3088: \end{equation}
3089: where
3090: \begin{equation*}
3091:     c(\gamma) \;:=\;
3092: \begin{cases}
3093: 1 & \ \text{if} \ \ \gamma \neq [1]\\
3094: \rule{0pt}{15pt}  \frac{ q([1]) - K(1)}{q([1])}  & \ \text{if} \ \ \gamma = [1]
3095: \end{cases} \;.
3096: \end{equation*}
3097: We are going to prove that~${\tt Z} (\widehat B) \ge 1$. Observe that setting $v_\ga := \Sigma_{[0],\ga}$ we can write
3098: \begin{equation*}
3099:     \Sigma_{\ga,\gb} = \Sigma_{[0],\gb} - \Sigma_{[0],\ga} = v_\gb - v_\ga\,.
3100: \end{equation*}
3101: Then we make a similarity transformation via the matrix~$L_{\ga,\gb} := \exp( v_\gb /2 ) \, \ind_{(\gb = \ga)}$, getting
3102: \begin{align*}
3103:     C_{\ga,\gb} &\;:=\; \big[ L \cdot \widehat B \cdot L^{-1} \big]_{\ga,\gb} \;=\; \exp \Big( \big( c(\gb - \ga) - 1 \big) \Sigma_{\ga,\gb} / 2 \Big) \cdot q(\gb - \ga)\\
3104:     & \;=\; \exp \Big( d \, \Sigma_{\ga,\ga + [1]} \, \ind_{(\gb - \ga = 1)} \Big) \cdot q(\gb - \ga) \,,
3105: \end{align*}
3106: where we have introduced the constant~$d := -K(1) / (\,2\, q([1])\,)$. Of course~${\tt Z}(\widehat B) = {\tt Z}(C)$. Also notice that by the very definition of~$\Sigma_{\ga,\gb}$ we have $\Sigma_{\ga,\ga + [1]} = \go^{(-1)}_{\ga + [1]} - \go^{(+1)}_{\ga + [1]}$, hence the hypothesis~$h_\go = 0$ yields~$\sum_{\ga \in \bbS} (\Sigma_{\ga,\ga + [1]}) = 0$.
3107: 
3108: Thus we are finally left with showing that ${\tt Z}(C) \ge 1$ where $C_{\ga,\gb}$ is an~$\bbS \times \bbS$ matrix of the form
3109: \begin{equation*}
3110:     C_{\ga,\gb} \;=\; \exp \big(\, w_\ga \, \ind_{(\gb - \ga = 1)} \,\big) \cdot q(\gb - \ga) \qquad \text{where} \qquad \sum_{\ga} w_{\ga} = 0 \qquad \sum_\gamma q(\gamma) = 1\,.
3111: \end{equation*}
3112: To this end, we introduce an interpolation matrix
3113: \begin{equation*}
3114:     C(t)_{\ga,\gb} \;:=\; \exp \big(\, t \cdot w_\ga \, \ind_{(\gb - \ga = 1)} \,\big) \cdot q(\gb - \ga) \,,
3115: \end{equation*}
3116: defined for $t \in \R$, and notice that $C(1) =C$. Let us denote by~$\eta(t):= {\tt Z}\big( C(t) \big)$ the Perron--Frobenius eigenvalue of~$C(t)$: as the entries of $C(t)$ are log--convex functions of~$t$, it follows that also $\eta(t)$ is log--convex, therefore in particular convex. Moreover $\eta(0)=1$ (the matrix~$C(0)$ is bistochastic) and using~\eqref{eq:der} one easily checks that $\frac{\dd}{\dd t} \eta(t) |_{t=0} = 0$.
3117: Since clearly $\eta(t) \ge 0$ for all~$t\in\R$, by convexity it follows that indeed $\eta(t) \ge 1$ for all~$t\in\R$, and the proof is complete.
3118: 
3119: