math0511561/cgg.tex
1: 
2: In this chapter we study the copolymer near a selective interface
3: model, defined in Section~\ref{sec:main_model}
4: of Chapter~\ref{ch:first}, in the {\sl random} case. We combine numerical computations
5: with rigorous arguments to get to a better understanding of the phase diagram
6: and of the path behavior. Our main aim is to provide evidences of the fact that
7: the critical line of the model~$h_c(\cdot)$ lies \emph{strictly} in between
8: the two bounds~$\underline h(\cdot)$ and~$\overline h(\cdot)$, defined in
9: \eqref{eq:sumupq} of Chapter~\ref{ch:first}, and to numerically analyze
10: the delocalization issues raised in \S~\ref{sec:pathwise} of Chapter~\ref{ch:first}.
11: A detailed outline of the results is given in~\S~\ref{sec:outline}.
12: 
13: \smallskip
14: The article~\cite{cf:CGG} has been taken from the content of this chapter.
15: 
16: \smallskip
17: \section{Introduction and results}
18: 
19: \smallskip
20: \subsection{Prelimiaries}
21: \label{sec:notation}
22: 
23: The notations we will use are those introduced in Section~\ref{sec:main_model}
24: of Chapter~\ref{ch:first}. We recall
25: in particular the definitions~\eqref{eq:Znew} and~\eqref{eq:pinned} of the
26: partition functions~$Z_{N,\go}^{\gl,h}$ and~$Z_{N,\go}^{\gl,h}(x)$ (we
27: will be mainly interested in the case~$x=0$). Also remember that for the
28: critical line~$h_c(\cdot)$ of our model we have the bounds
29: \begin{equation} \label{eq:sumup_alt}
30:     \underline h (\gl) \;\le\; h_c (\gl) \;\le\; \overline{h} (\gl),
31: \end{equation}
32: see Theorem~\ref{th:sumup} of Chapter~\ref{ch:first}. For what follows we set
33: \begin{equation}
34: \label{eq:hm}
35: h^{(m)}(\gl) \, = \, \frac 1{2m\gl } \log \M\left( -2m\gl \right),
36: \end{equation}
37: for $m >0$, where we recall that $\M (\ga) := \bbE[\exp (\ga \go_1)]$. Observe that the curves $\underline h(\cdot)$ and $\overline h(\cdot)$
38: correspond respectively to $m=2/3$ and $m=1$, and that for every~$m$ we have
39: $\frac{\dd}{\dd\gl} h^{(m)}(\gl) |_{\gl=0} = m$.
40: 
41: \medskip
42: 
43: Before proceeding, we present a different viewpoint on the process:
44: this turns out to be useful for the intuition and it will be used
45: in some technical steps. We call  $\eta$  the first
46: return time of the walk $S$ to $0$, that is $\eta:=\inf\left\{  n\ge
47: 1:S_{n}=0\right\} $, and  set $K(2n):=\bP \left(\eta=2n\right)$ for $n\in\N$.
48: It is well known that $K(\cdot)$ is decreasing on the even natural numbers and
49: \begin{equation}
50: \label{eq:asympt}
51: \lim_{x\in 2\N, x\to \infty} x^{3/2} K(x)=
52: \sqrt{2/\pi },
53: %=:c_K,
54: \end{equation}
55: see e.g. \cite[Ch. 3]{cf:Feller}.
56: Let the IID sequence $\left\{ \eta_j \right\}_{j=1,2, \ldots}$
57: denote the inter--arrival times at $0$ for $S$, and we set $\tau_k := \eta_0 + \ldots + \eta_k$.
58:  If we introduce also
59:  $\ell _N =\max\{j \in \N \cup \{0\} :
60: \tau_j \le N\}$, then by  exploiting
61: the up--down symmetry of the excursions of $S$  we directly obtain
62: \begin{equation}
63: \begin{split}
64: \label{eq:reducttoexc}
65: & Z_{N,\go} (0) \, =\, \bE \left[ \prod_{j=1}^{\ell_N} \varphi \Bigg(\gl \sum_{n=\tau_{j-1}+1}^{\tau_j} \go_n + \gl h  \eta_j \Bigg) ; \tau_{\ell_N} = N \right] \\
66:     & \quad =\, \sum_{l=0}^N \sumtwo{x_0, \ldots, x_l \in 2\N}{0=:x_0 < \ldots < x_l:= N} \prod_{i=1}^l \; \varphi \Bigg(\gl \sum_{n=x_{i-1}+1}^{x_i} \go_n + \gl h  (x_i - x_{i-1}) \Bigg) \; K(x_i - x_{i-1}) \,,
67: \end{split}
68: \end{equation}
69: with $\varphi (t) : =   \left( 1+\exp(-2t)\right) / 2$.
70: Of course the formula for $Z_{N,\go}$
71: is just  slightly different.
72: 
73: Formula \eqref{eq:reducttoexc} reflects the fact that what really
74: matters for the copolymer are the return times to the interface.
75: 
76: 
77: 
78: \smallskip
79: \subsection{Outline of the results}
80: \label{sec:outline}
81: 
82: 
83: Formula \eqref{eq:sumup_alt} leaves an important gap, that
84: hides the only partial understanding of the nature of this delocalization/localization
85: transition.
86: Our purpose is to go toward filling this gap: our results are both of theoretical and numerical nature.
87: At the same time we address the delocalization issues raised
88: in \S~\ref{sec:pathwise} of Chapter~\ref{ch:first}, which are intimately related with the precise
89: asymptotic behavior of $Z_{N,\go}$ and of  $Z_{N,\go}(0)$.
90: More precisely:
91: 
92: \smallskip
93: \begin{enumerate}
94: \item In Section~\ref{sec:testing} we present a statistical test with explicit error bounds, see Proposition~\ref{th:stat}, based on
95: super--additivity and concentration inequalities, to state that a point $(\gl, h)$ is localized. We apply this test to show that, with a very low level of error, the lower bound $h=\underline h(\gl)$ does not coincide with the critical line.
96: \item \rule{0pt}{14pt}In Section~\ref{sec:lb} we give the outline of a new proof of the lower bound
97: $h_c(\cdot) \ge \underline h (\cdot)$. The details of the proof are in \S~\ref{app:prooflb} and we point out
98: in particular Proposition~\ref{prop:stopping1}, that gives a necessary and sufficient
99: condition for localization. This viewpoint on the transition, derived from
100: \cite[\S~4]{cf:GT}, helps substantially in interpreting the {\sl irregularities} in the
101: behavior  of $\left\{ Z_{N, \go}\right\}_N$ as $N \nearrow \infty$.
102: \item \rule{0pt}{14pt}In Section~\ref{sec:path} we pick up the conjecture of Brownian scaling
103: in the delocalized regime both in the intent of testing it and
104: in trying to asses with reasonable confidence that $(\gl, h)$ is in the interior
105: of $\cD$. In particular, we present quantitative evidences in favor of the fact that the upper bound $h= \overline h (\gl)$ is strictly greater than the critical line. We stress that this is a very delicate issue, since delocalization, unlike
106: localization, does not appear to be reducible to a finite volume issue.
107: \item \rule{0pt}{14pt}Finally, in Section~\ref{sec:guess}, we report the results of a numerical attempt
108: to determine the critical curve. While this issue has to be treated with care,
109: mostly for the reasons raised in point 4 above, we observe
110: a surprising phenomenon: the critical curve appears to be  very close
111: to $h^{(m)}(\cdot)$ for a suitable value of $m$. By the universality
112: result proven in \cite{cf:GT}, building on the free energy Brownian scaling result proven
113: in \cite{cf:BdH}, the slope at the origin of $h_c(\cdot)$ does not depend on the law of $\go$.
114: Therefore if really $h^{(m)}(\cdot)= h_c (\cdot)$, since the slope at the origin
115: of $h^{(m)}(\cdot)$ is $m$,
116: $m$ is the  universal constant we are looking for.
117: We do not believe that the numerical evidence
118: allows to make a clear cut  statement, but what we observe is compatible
119: with such a possibility.
120: \end{enumerate}
121: \smallskip
122: 
123: We point out that our numerical results are based on a numerical computation
124: of the partition function $Z_{N,\go}$, exploiting the standard transfer--matrix
125: approach (this item is discussed in more details in \S~\ref{app:algo}).
126: 
127: 
128: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
129: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
130: 
131: 
132: \smallskip
133: \section{A statistical test for the localized phase}
134: \label{sec:testing}
135: 
136: \smallskip
137: \subsection{Checking localization at finite volume}
138: \label{sec:superadd}
139: At an intuitive level one is led to believe that, when the copolymer is localized,
140: it should be possible to detect it
141: by looking at the system before the infinite volume limit.
142: This intuition is due to the fact that in the localized phase
143: the length of each excursion is finite, therefore for $N$ {\sl much larger }
144: that the {\sl typical} excursion length one should already observe the
145: localization phenomenon in a quantitative way.
146: The system being disordered of course does not help,
147: because it is more delicate to make sense of what
148: {\sl typicality} means in a non translation invariant set--up.
149:  However the translation invariance can be recovered
150:  by averaging and in fact it turns out to be rather easy to give
151:  a precise meaning to the intuitive idea we have just mentioned.
152:  The key word here is super--additivity of the averaged free energy.
153:  \smallskip
154: 
155:  In fact by considering only the $S$ trajectories such that $S_{2N}=0$ and by applying
156:  the Markov property of $S$ one directly verifies
157:   that for any $N, M\in \N$
158:  \begin{equation}
159:  \label{eq:superadd}
160:  Z_{2N+2M, \go} (0)\, \ge \, Z_{2N, \go}(0)\, Z_{2M, \theta^{2N} \go}(0),
161:  \end{equation}
162:  $(\theta \go)_n=\go_{n+1}$,
163:  and therefore
164:  \begin{equation}
165:  \left\{ \, \bbE\log Z_{2N, \go} (0)
166:  \right\}_{N=1,2, \ldots}
167:   \end{equation}
168:   is a super--additive sequence, which immediately entails
169:   the existence of the limit of $ \bbE[\log Z_{2N, \go}(0)]/2N$ and the fact that
170:   this limit coincides with the supremum of the sequence.
171:   Therefore from the existence of the quenched free energy we have that
172:   \begin{equation}
173:   \tf(\gl, h) \, =\, \sup_N \frac 1{2N} \bbE\log Z_{2N, \go}(0)\,.
174:   \end{equation}
175: In a more suggestive way one may say that:
176: \begin{equation}
177: \label{eq:loc_char}
178: (\gl, h) \in \cL \  \  \Longleftrightarrow  \ \
179: \text{there exists } N\in \N \ \, \text{such that} \,  \  \bbE\log Z_{2N, \go}(0)>0\,.
180: \end{equation}
181: The price one pays for working with a disordered system
182: is precisely in taking the $\bbP$--expectation
183: and from the numerical viewpoint it is an heavy price:
184: even with the most positive attitude one cannot expect
185: to have access to  $\bbE\log Z_{2N, \go}(0)$
186: by direct numerical computation for $N$ above $10$.
187: Of course in principle small values of $N$ may suffice
188: (and they do in some cases, see Remark~\ref{rem:N=2}), but they do not
189: suffice to tackle the specific issue we are interested in.
190: We elaborate at length on this interesting issue in \S~\ref{sec:computer_assisted}.
191: \smallskip
192: 
193: \begin{rem}
194: \label{rem:N=2}
195: \rm
196: An elementary application of the localization criterion \eqref{eq:loc_char} is
197: obtained for $N=1$: $(\gl, h)\in \cL$ if
198: \begin{equation}
199: \label{eq:N=2}
200: \bbE\left[
201: \log\left(
202: \frac 12+ \frac 12 \exp\left(
203: -2\gl \left(\go_1+ \go_2+2h\right)
204: \right)
205: \right)
206: \right]>0.
207: \end{equation}
208: In the case $\bbP (\go_1=\pm 1)=1/2$ from
209: \eqref{eq:N=2}
210: we obtain that for $\gl$ sufficiently large
211: $h_c(\gl) > 1- c/\gl$, with $c= (1/4)\log(2\exp(4) -1)\approx1.17$.
212: From $\underline{h}(\cdot)$ we obtain the same type of bound,
213: with $c=(3/4)\log 2\approx 0.52$. This may raise some hope that
214: for $\gl$ large an explicit, possibly computer assisted, computation
215: for small values of $N$ of $\bbE \log Z_{2N, \go}(0)$ could
216: lead to new estimates. This is not the case, as we show in~\S~\ref{sec:computer_assisted}.
217: \end{rem}
218: 
219: \smallskip
220: \subsection{Testing by using concentration}
221: 
222: In order to decide whether  $\bbE\log Z_{2N, \go}(0)>0$ we
223: resort to a Montecarlo evaluation of $\bbE\log Z_{2N, \go}(0)$
224: that can be cast into a statistical test with explicit error bound
225: by means of concentration of measure ideas.
226: This procedure is absolutely general, but we have to choose
227: a set--up for the computations and we take the simplest:
228:  $\bbP(\go_1=+1)=\bbP(\go_1=-1)=1/2$.
229:  The reason for this choice is twofold:
230:  \begin{itemize}
231:  \item if $\go_1$ is a bounded random variable, a Gaussian concentration
232:  inequality holds and if $\go$ is symmetric and it takes only two values
233:  then one can improve on the explicit constant in such an inequality.
234:  This speeds up in a non negligible   way the computations;
235:  \item generating {\sl true randomness}
236: is out of reach, but playing head and tail is certainly
237: the most elementary case in such a far reaching task (the random numbers
238: issue is briefly discussed in \S~\ref{app:algo} too).
239:  \end{itemize}
240: 
241: \smallskip
242: A third reason to restrict testing to the Bernoulli case is explained at the end of the
243: caption of Table \ref{tbl:2}.
244: \smallskip
245: 
246: We start the testing procedure by stating the null hypothesis:
247: \begin{equation}
248: \label{eq:H0}
249: \text{H}0: \  \ \bbE\log Z_{2N, \go}(0)\le 0.
250: \end{equation}
251: $N$ in H0  can be chosen
252: arbitrarily. We stress that refusing H0 implies $\bbE\log Z_{2N, \go}(0)>0$,
253: which by \eqref{eq:loc_char} implies localization.
254: 
255: The following concentration inequality for Lipschitz functions holds
256: for the uniform measure on $\{-1,+1\}^N$: for every
257: function $G_N:\{-1,+1\}^N \to \R$
258: such that $\vert G_N(\go)-G_N(\go^\prime )\vert
259: \le C_{\text{Lip}} \sqrt(\sum_{n=1}^N (\go_n-\go^\prime_n)^2)$, where
260: $C_{\text{Lip}}$ a positive constant
261: and $G_N(\go)$ is an abuse of notation for $G_N(\go_1, \ldots, \go_N)$, one has
262: \begin{equation}
263: \label{eq:concentration}
264: \bbE\left[\exp\left(\ga \left(G_N(\go)- \bbE[G_N(\go)]\right)\right)\right]\, \le\,
265: \exp\left(\ga^2C_{\text{Lip}}^2\right),
266: \end{equation}
267: for every $\ga$.
268: Inequality \eqref{eq:concentration} with an extra factor $4$ at the exponent
269: can be extracted from the proof of Theorem 5.9, page 100 in \cite{cf:Ledoux}.
270: Such an inequality holds for variables taking values in $[-1,1]$:
271: the factor $4$ can be removed for the particular case we are considering
272: (see \cite[p. 110--111]{cf:Ledoux}).
273: In our case $G_N(\go)= \log Z_{2N,\go}(0)$. By applying the
274:  Cauchy--Schwarz inequality one obtains that $G_N$ is Lipschitz with
275: $C_{\text{Lip}}= 2\gl\sqrt{N}$.
276: Let us now consider an IID sequence  $\{ G^{(i)}_N(\go)\}_i$
277: with $G^{(1)}_N (\go)=G_N(\go)$: if H0 holds then we have that for every $n\in \N$, $u>0$
278: and $\ga= un/8\gl^2N$
279: \begin{equation}
280: \begin{split}
281: \bbP\left(
282: \frac 1n \sum_{i=1}^n G^{(i)}_N(\go) \ge u
283: \right)
284: \, &\le  \,
285: \bbE\left[
286: \exp\left(
287: \frac\ga n \left(G_N(\go)- \bbE[G_N(\go)]\right)
288: \right)
289: \right]^n
290: \exp\left(-\ga\left( u- \bbE[G_N(\go)]\right)\right)
291: \\
292: &\le \,
293: \exp\left(\frac{4 \ga^2 \gl^2N}{n}-\ga u\right)
294: \\
295: &= \, \exp\left( -\frac{u^2 n}{16 \gl^2 N}\right) .
296: \end{split}
297: \end{equation}
298: 
299: 
300: Let us sum up what we have obtained:
301: 
302: \medskip
303: \begin{proposition}
304: \label{th:stat}
305: Let us call $\widehat u_n$ the average of a sample of $n$ independent
306: realizations of $\log Z_{2N,\go}^{\gl,h}(0)$. If $\widehat u_n>0$ then we may
307: refuse H0, and therefore  $(\gl,h)\in \cL$, with a level
308: of error not larger than $\exp\left( -{\widehat u_n ^2 n}/{16 \gl^2 N}\right)$.
309: \end{proposition}
310: \medskip
311: 
312: 
313: 
314: 
315: 
316: 
317: 
318: \smallskip
319: \subsection{Numerical tests}
320: We report in Table \ref{tbl:1}
321:  the most  straightforward application
322: of Proposition \ref{th:stat}, obtained by a numerical computation of $\log Z_N$ for a sample of~$n$ independent environments~$\go$.
323: We aim
324: at seeing how far above $\underline{h}(\cdot)$ one can go
325: and still claim localization,
326: keeping a reasonably small probability of error.
327: 
328: \begin{table}[h]
329: \begin{center}
330: \begin{tabular}{|c|c|c|c|}
331: \hline
332: $\gl $ &$0.3$&$0.6$&$1$\\
333: \hline
334: $h$&  0.22 &0.41&0.58\\
335: \hline
336: $p$--value& $1.5\times 10^{-6}$&$9.5\times 10^{-3}$&$1.6 \times 10^{-5}$\\
337: \hline
338: $\underline{h}(\gl)$& 0.195 &0.363&0.530\\
339: \hline
340: $\overline{h}(\gl)$
341:  & 0.286 &0.495&0.662\\
342: \hline
343: $N$& 300000 &500000&160000\\
344: \hline
345: $n$& 225000 &330000&970000\\
346: \hline
347: C. I.  99\% & $7.179\pm0.050$ &$9.011\pm 0.045$& $7.643 \pm 0.025$\\
348: \hline
349: \end{tabular}
350: \end{center}
351: \bigskip
352: \caption{\label{tbl:1}
353: According to our numerical computations,
354: the three pairs $(\gl,h)$ are in $\cL$ and this has been tested with the stated
355:  $p$--values (or probability/level of error).
356: We report the values of $\overline{h}(\gl)$ and $\underline{h}(\gl)$ for reference.
357: Of course in these tests there is quite a bit of freedom in the choice of $n$
358: and $N$: notice that $N$ enters in the evaluation of the $p$--value also
359: because a larger value of $N$ yields a larger value of $\bbE \log Z_{2N,\go}^{\gl,h}(0)$.
360: In the last line we report  standard Gaussian $99\%$ confidence intervals for
361: $\bbE \log Z_{2N,\go}^{\gl,h}(0)$. Of course the $p$--value under the Gaussian assumption
362: turns out to be totally negligible.}
363: \end{table}
364: 
365: \begin{rem}
366: \rm
367: One might be tempted to interpolate between the values in Table  \ref{tbl:1},
368: or possibly to get results for small  values of $\gl$ in order to
369: extend the result of the test to the slope of the critical curve in the origin.
370: However the fact that $h_c(\gl)$ is strictly increasing does not help much in this direction
371: and the same is true for
372:  the finer result, proven in \cite{cf:BG}, that $h_c(\gl)$ can be written as $U(\gl)/\gl$,
373: $U(\cdot)$ a convex function.
374: \end{rem}
375: 
376: 
377: \smallskip
378: \subsection{Improving on $\underline{h}(\cdot)$ is uniformly hard}
379: \label{sec:computer_assisted}
380: One can get much smaller $p$--values at little computational cost  by choosing
381: $h$ {\sl just above} $\underline{h}(\gl)$. As a matter of fact
382: a natural choice is for example  $h=h ^{(0.67)}(\gl)>\underline{h}(\gl)$, recall
383: \eqref{eq:hm}, for a set of values of $\gl$, and this is part of the content of
384: Table \ref{tbl:2}: in particular $\bbE \log Z_{2N_+,\go}^{\gl,h^{(0.67)}(\gl)}(0)  >0$
385: with a probability of error smaller than $10^{-5}$ for the values of $\gl$
386: between $0.1$ and $1$. However we stress that
387: for some of these $\gl$'s we have a much smaller $p$--value, see the caption
388: of Table \ref{tbl:2}, and that the content of this table is much richer
389: and it approaches also the question of whether or not
390: a symbolic computation  or some other form of
391: computer assisted argument could lead to $h_c(\gl)>\underline{h}(\gl)$ for
392: some $\gl$, and therefore for $\gl$ in an interval. Since such an argument would require
393: $N$ to be {\sl small}, intuitively the hope resides in large values of $\gl$, recall also
394:   Remark~\ref{rem:N=2}. It turns out that one needs in any case
395:   $N$ larger than $700$ in order to observe a localization phenomenon
396:   at $h^{(0.67)}(\gl)$.
397: We now give some details on the procedure that leads to Table  \ref{tbl:2}.
398: 
399: 
400: \begin{table}[h]
401: \begin{center}
402: \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
403: \hline
404: $\gl $ &$0.05 (\star)$&$0.1$&$0.2$&$0.4$&$0.6$&$1$&$2(\star)$&$4(\star\star )$&$8(\star\star )$\\
405: \hline
406: $N_+$&750000&190000&40000&9500&4250&1800&900&800&800\\
407: \hline
408: $N_-$&600000&130000&33000&7500&3650&1550&750&700&700\\
409: \hline
410: \end{tabular}
411: \end{center}
412: \bigskip
413: \caption{\label{tbl:2}
414: For a given $\gl$, both $\bbE \log Z_{2N_+,\go}^{\gl,h^{(0.67)}(\gl)}(0)  >0$ and
415: $\bbE \log Z_{2N_-,\go}^{\gl,h^{(0.67)}(\gl)}(0)  <0$ with a probability of error
416: smaller than $10^{-5}$ (and in some cases much smaller than that).
417: Instead for the two cases marked by a $(\star)$ the level of error
418: is rather between $10^{-2}$ and  $10^{-3}$. For large values of $\gl$, the two cases marked with
419:  $(\star \star)$, it becomes computationally
420: expensive to reach small $p$--values. However, above $\gl=3$ one observes
421: that the values of $Z_{2N,\go}(0)$
422: essentially do not depend anymore on the value of $\gl$. This can be interpreted
423: in terms of convergence to a limit ($\gl \to \infty$) model, as it is explained in Remark~\ref{rem:starstar}.
424: If we then make the hypothesis that this limit model sharply describes the
425: copolymer along the curve $(\gl, h^{(m)}(\gl))$ for $\gl$ sufficiently large and we apply the concentration
426: inequality, then the given values of $N_+$  and $N_-$ are tested
427: with a very small probability of error. Since the details of such a procedure are
428: quite lengthy we do not report them here.
429: %\\
430: We have constructed (partial) tables also for different laws of $\go$,
431: notably $\go_1 \sim N(0,1)$,  and they turned out
432: to yield larger, at times substantially larger, values of $N_\pm (\gl)$. %{\bf For internal use}: Gauss $\gl=.4$ $[12000,14000]$, $\gl=.6$ $[7000,8000]$, $\gl=1$ $[10000,12000]$, $\gl=1.2$ $[18000,20000]$.
433: }
434: \end{table}
435: 
436: First and foremost, the concentration argument
437: that leads to Proposition~\ref{th:stat} is symmetric
438: and it works for deviations below the mean as well as above.
439: So we can, in the very same way, test the null hypothesis
440: $\bbE\log Z_{2N, \go}(0)> 0$ and, possibly,  refuse it
441: if $\hat u _n <0$, exactly with the same $p$--value
442: as in Proposition~\ref{th:stat}.
443: Of course an important part of Proposition~\ref{th:stat} was coming
444: from the finite volume localization condition \eqref{eq:loc_char}:
445:  we do not have an analogous statement for delocalization
446: (and we do not expect that there exists one).
447: But, even if $\bbE\log Z_{2N, \go}(0)\le 0$ does not imply delocalization,
448: it says at least that it is pointless to try to prove localization
449: by looking at a system of that size.
450: 
451: In Table~\ref{tbl:2} we show
452: two values of the system size $N$,  $N_+$ and $N_-$,
453: for which, at a given $\gl$, one has that
454: $\bbE\log Z_{2N_+, \go}(0)> 0$ and  $\bbE\log Z_{2N_-, \go}(0)< 0$
455: with a fixed probability of error (specified in the caption of the Table).
456: It is then reasonable to guess that the transition from
457: negative to positive values of $\bbE\log Z_{\cdot , \go}(0)$ happens
458: for $N\in (N_-,N_+)$. There is no reason whastoever to expect that
459:  $\bbE\log Z_{N, \go}(0)$ should be monotonic in $N$ but according to our numerical result it is not unreasonable to expect that monotonicity should set in for $N$ large or, at least, that for $N< N_-$ (respectively $N>N_+$)
460:  $\bbE\log Z_{2N, \go}(0)$ is definitely negative
461: (respectively positive).
462: 
463: \begin{figure}[h]
464: \begin{center}
465: \leavevmode
466: \epsfysize =8 cm
467: %\epsfxsize =10 cm
468: \psfragscanon
469: \psfrag{N}[c][l]{$N$}
470: \psfrag{lambda}[c][l]{ $\gl$}
471: \epsfbox{nca.eps}
472: \end{center}
473: \caption{\label{fig:NCA} A graphical representation  of
474: Table~\ref{tbl:2}. The plot is log--log, and a $\gl^{-c}$ behavior is rather
475: evident, $c$ is about $2.08$. This can be nicely interpreted in terms
476: of the coarse graining technique in the proof of the weak interaction scaling
477: limit of the free energy in \cite{cf:BdH}: from that argument one extracts that
478: if $\gl $ is small the excursions that give a contribution to the free
479: energy have {\sl typical} length $\gl^{-2}$ and that
480: in the limit the polymer is just made up by this type of excursions.
481: One therefore  expects
482: that it suffices a system of size $N(\gl)$, with $\lim_{\gl \searrow 0} \gl^2 N(\gl)=+\infty$,
483: to observe localization if $m<h_c^\prime( 0)$, $h=h^{(m)}(\gl)= m \gl (1+o(1))$ and $\gl$ is small.}
484: \end{figure}
485: 
486: 
487: \begin{rem}
488: \label{rem:starstar}
489: \rm
490: As pointed out in the caption of Table  \ref{tbl:2}, from numerics one observes
491: a very sharp convergence to a $\gl$ independent behavior as $\gl$
492: becomes large, along the line $h=h^{(m)}(\gl)$.
493: This is easily interpreted if one observes that $h^{(m)}(\gl)= 1-((\log 2) /2m\gl)+O(\exp(-4m\gl))$
494: so that
495: \begin{equation}
496: \label{eq:lim_mod}
497: \lim_{\gl \to \infty} \exp\left(-2 \gl \sum_{n=1}^N
498: \left( \go_n +h\right) \Delta_n \right)\, =\, \exp\left(\frac{\log 2}{m}\sum_{n=1}^N \Delta_n\right)
499: \ind_{\left\{\sum_{n=1}^N \Delta_n (1+\go_n)=0\right\} } (S).
500: \end{equation}
501: This corresponds to the model where a positive charge never enters the lower half-plane and where the energy of a configuration is proportional to the number of negative charges in the lower half-plane.
502: \end{rem}
503: 
504: %\begin{rem}
505: %\label{rem:table1}\rm
506: %In the light of this section we can probably better interpret the values in Table  \ref{tbl:1}. In fact, with reference to \eqref{eq:hm}, the values of $h$ chosen  for $\gl=0.6$ and $1$ correspond to $m$ slightly  larger than $0.77$, while in the case of $\gl=0.3$ the value of $h$ corresponds to $m$ slightly smaller than $0.76$.
507: %\end{rem}
508: 
509: 
510: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
511: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
512: 
513: \smallskip
514: \section{Lower bound strategies versus the true strategy}
515: \label{sec:lb}
516: 
517: \subsection{An approach to lower bounds on the critical curve}
518: \label{sec:lb_outline}
519: In this section we give an outline of a new derivation of the lower bound
520: \begin{equation}
521: \label{eq:mainBG}
522: \underline h (\gl)\le h_c(\gl),
523: \end{equation}
524: with $\underline h (\gl)$ defined in \eqref{eq:sumupq} of Chapter~\ref{ch:first}.
525: The complete proof may be found in \S~\ref{app:prooflb}.
526: The argument takes inspiration from the ideas used in the proof of Proposition 3.1 in \cite{cf:GT} and, even if it is essentially the proof of \cite{cf:BG} in disguise,
527: in the sense that the selection of the random walk trajectories that are kept
528: and whose energy contribution is evaluated
529: does not differ too much (in a word: the {\sl strategy} of the polymer
530: is similar), it is however conceptually somewhat different and it
531: will naturally lead to some considerations on the precise
532: asymptotic behavior of $Z_{N, \go}$ in the delocalized phase and
533: even in the localized phase close to criticality.
534: 
535: \bigskip
536: 
537: The first step in our proof of (\ref{eq:mainBG}) is a different way of looking at localization. For any fixed positive number $C$ we introduce the stopping time (with respect to the natural filtration of the sequence $\{\go_n\}$) $T^C = T^{C,\gl,h}(\go)$ defined by
538: \begin{equation}
539: \label{eq:T^C}
540: T^{C,\gl,h}(\go) := \inf \{N\in 2\N:\ Z_{N,\go}^{\gl,h}(0) \ge C\}\,.
541: \end{equation}
542: 
543: The key observation is that if $\bbE[T^C] < \infty$ for some $C>1$, then the polymer is localized. Let us sketch a proof of this fact (for the details, see Proposition~\ref{prop:stopping1} of \S~\ref{app:prooflb}): notice that by the very definition of $T^C$ we have $Z_{T^C(\go),\go}(0) \ge C$. Now the polymer that is in zero at $T^C(\go)$ is equivalent to the original polymer, with a translated environment $\go'=\theta^{T^C} \go$, and setting $T_2(\go) := T^C (\go')$ we easily get $Z_{T_1(\go)+T_2(\go),\go}(0) \ge C^2$ (we have put $T_1(\go):=T^C(\go)$). Notice that the new environment $\go'$ is still typical, since $T^C$ is a stopping time, so that $T_2$ is independent of $T_1$ and has the same law. This procedure can be clearly iterated, yielding an IID sequence $\{T_i(\go)\}_{i=1,2,\ldots}$ that gives the following lower bound on the partition function:
544: \begin{equation} \label{eq:lowbound}
545: Z_{T_1(\go)+\ldots +T_n(\go),\go}(0) \ge  C^n\,.
546: \end{equation}
547: From this bound one easily obtains that
548: \begin{equation}
549: \label{eq:forappB}
550: \tf(\gl,h) \stackrel{\text{a.s.}}{= } \lim_{n\to\infty} \frac{\log Z_{T_1(\go)+\ldots +T_n(\go),\go}(0)}{T_1(\go)+\ldots +T_n(\go)} \ge \frac{\log C}{\bbE[T^C]}\,,
551: \end{equation}
552: where we have applied the strong law of large numbers, and localization follows since by hypothesis $C>1$ and $\bbE[T^C]<\infty$.
553: 
554: \medskip
555: 
556: \begin{rem}
557: \label{rem:reciprocal}
558: \rm
559: It turns out that also the reciprocal of the claim just proved holds true, that is \textsl{the polymer is localized if and only if $\bbE[T^C]<\infty$}, with an arbitrary choice of $C>1$, see Proposition~\ref{prop:stopping1}. In fact the case $\bbE[T^C]=\infty$ may arise in two different ways:
560: \begin{enumerate}
561: \item the variable $T^C$ is defective, $\bbP[T^C=\infty]>0$: in this case with positive probability $\{Z_{N,\go}(0)\}_N$ is a bounded sequence, and delocalization follows immediately;
562: \item\label{en:scenario} the variable $T^C$ is proper with infinite mean, $\bbP[T^C=\infty]=0,\ \bbE[T^C]=\infty$: in this case we can still build a sequence $\{T_i(\go)\}_{i=1,2,\ldots}$ defined as above and this time the lower bound \eqref{eq:lowbound} has \textsl{subexponential} growth. Moreover it can be shown that in this case the lower bound \eqref{eq:lowbound} gives the true free energy, cf. Lemma~\ref{lem:chop}, which therefore is zero, so that delocalization follows also in this case.
563: \end{enumerate}
564: As a matter of fact, it is highly probable that in the interior of the delocalized phase
565: $Z_{N,\go}(0)$ vanishes $\bbP (\dd \go)$--a.s.
566: when $N \to \infty$ and this would rule out the scenario (\ref{en:scenario}) above, saying that for $C>1$ the random variable $T^C$ must be either integrable or defective. We take up again this point in Sections~\ref{sec:path} and~\ref{sec:guess}: we feel that this issue
567: is quite crucial in order to fully understand the delocalized phase
568: of disordered models.
569: \end{rem}
570: 
571: %\smallskip
572: 
573: \begin{rem}
574: \label{rem:analogy}
575: \rm
576: Dealing directly with $T^C$ may be difficult.
577: Notice however that if one finds
578:  a random time (by this we mean simply an integer--valued random variable)
579:  $T=T(\go)$ such that
580:  \begin{equation} \label{eq:but}
581: Z_{T(\go), \go}(0)\ge C>1\,, \qquad \text{with \ } \bbE[T]<\infty\,,
582: \end{equation}
583: then localization follows. This is simply because
584: this implies $T^C \le T$ and hence $\bbE[T^C]<\infty$. Therefore localization is equivalent to the condition $\log Z_{T(\go), \go}(0)>0$ for an \textsl{integrable} random time $T(\go)$: we would like to stress the analogy between this and the criterion for localization given in \S~\ref{sec:superadd}, see \eqref{eq:loc_char}.
585: \end{rem}
586: 
587: 
588: 
589: \medskip
590: 
591: Now we can turn to the core of our proof: we are going to show that for every $(\gl,h)$ with $h<\underline h (\gl)$ we can build a random time $T=T(\go)$ that satisfies \eqref{eq:but}.
592: The construction of $T$
593:  is based on the  idea that for $h>0$ if localization prevails
594:  is because of rare $\go$--stretches that invite the polymer
595:  to spend time in the lower half--plane in spite of the action of $h$.
596: 
597: The strategy we use consists in looking
598: for $q$--atypical stretches of length at least $M\in 2\N$, where $q<-h$ is
599: the average charge of the stretch. Rephrased a bit more precisely,
600: we are looking for the smallest $n\in 2\N$ such that
601: $\sum_{i=n-k+1}^n \go_i /k <q$ for some even integer $k \ge M$.
602: It is well known that such a random variable grows, in the sense
603: of Laplace, as $\exp(\Sigma(q) M)$ for $M\to \infty$, where $\gS(q)$ is the Cramer functional
604: \begin{equation}
605: \label{eq:cramer}
606: \gS(q) := \sup_{\ga\in\R} \{\ga q-\log \M(\ga)\}\,.
607: \end{equation}
608: One can also show without much effort that the length of such a stretch
609: cannot be much longer than $M$.
610: Otherwise stated, this is the familiar statement that the longest $q$--atypical
611: sub--stretch of $\go_1, \ldots , \go_N$ is of typical length $\sim \log N/\Sigma (q)$.
612:  So $T(\go)$ is for us the end--point
613: of a $q$--atypical stretch of length approximately $(\log T(\go))/\Sigma(q)$:
614: by looking for sufficiently long $q$--atypical stretches
615: we have always the freedom to choose $T(\go) \gg 1$, in such a
616: way that also $\log T(\go) \ll T(\go)$ and this is helpful for the estimates.
617: So let us bound $Z_{T(\go), \go}$ from
618: below by considering only the trajectories of the walk that
619: stay in the upper half--plane up to the beginning of the $q$--atypical stretch and
620: that are negative in the stretch, coming back to zero at step $T(\go)$
621: (see Fig.~\ref{fig:ZT}: the polymer is cut at the first dashed vertical line).
622: The contribution of these trajectories is easily evaluated: it is approximately
623: \begin{equation}
624: \label{eq:int1_cgg}
625: \left(
626: \frac {1} {T(\go)^{3/2}}
627: \right)
628:  \exp\left( -2\gl (q+h)  \frac{\log T(\go)}{\Sigma(q)}\right).
629: \end{equation}
630: For such an estimate we have used
631: \eqref{eq:asympt} and $\log T(\go) \ll T(\go)$
632: both in writing the probability that the first return to zero
633: of the walk is at the beginning of the $q$--atypical stretch and in neglecting the probability
634: that the walk is negative inside the stretch.
635: It is straightforward to see that if
636: \begin{equation}
637: \label{eq:int2_cgg}
638: \frac {4\gl} 3 h < -\frac {4\gl} 3 q -  \Sigma(q),
639:  \end{equation}
640: and if $T(\go)$ is large, then also  the quantity in
641:  \eqref{eq:int1_cgg} is large. We can still optimize this procedure by choosing $q$ (which must be sufficently negative, i.e. $q < -h$).
642:  By playing with
643: \eqref{eq:cramer}
644: one sees that one can choose $q_0\in \R$ such that for $q=q_0$ the right--hand side
645: in \eqref{eq:int2_cgg} equals $\log \M (-4\gl/3)$ and
646: if $h<\log \M (-4\gl/3)/(4\gl/3) = \underline{h} (\gl)$
647:  then $q_0 <-h$. This argument therefore is saying that there
648:  exists $C>1$ such that
649:  \begin{equation}
650:  \label{eq:C1}
651:  Z_{T(\go), \go}(0) \, \ge \, C,
652: \end{equation}
653: for every $\go$. It only remains to show that $\bbE[T]<\infty$: this fact, together with a detailed proof of the argument just presented can be found in \S~\ref{app:prooflb}.
654: 
655: 
656: 
657: \medskip
658: 
659: \begin{figure}[h]
660: \begin{center}
661: \leavevmode
662: \epsfysize =5 cm
663: %\epsfxsize =10 cm
664: \psfragscanon
665: \psfrag{0}[c][l]{$0$}
666: \psfrag{l}[c][l]{ $\ell$}
667: \psfrag{n}[c][l]{ $n$}
668: \psfrag{h}[l][l]{ $L$}
669: \psfrag{S}[c][l]{$S$}
670: \psfrag{T}[c][l]{$T(\go)$}
671: \epsfbox{ZTom.eps}
672: \end{center}
673: \caption{\label{fig:ZT}
674: Inequality \eqref{eq:beyondT}
675: comes simply from restricting the evaluation of $Z_{T(\go) +L, \go} $ to the trajectories
676: visiting the {\sl $q$--atypical} stretch of length $\ell$  and by staying away from
677: the unfavorable solvent after that.
678: }
679: \end{figure}
680: 
681: \begin{figure}[h]
682: \begin{center}
683: \leavevmode
684: \epsfysize =14 cm
685: %\epsfxsize =10 cm
686: \psfragscanon
687: \psfrag{x}[c][l]{\small $N$}
688: \psfrag{y}[c][l]{\small $\log  Z_{2N, \go}^{\gl, h}$}
689: \psfrag{a}[l][l]{A: $h=0.42$}
690: \psfrag{b}[l][l]{B:  $h=0.44$}
691: \psfrag{c}[l][l]{C:  $h=0.43$}
692: \psfrag{d}[l][l]{D:  $h=0.43$ (Zoom)}
693: \epsfbox{for4.eps}
694: \end{center}
695: \caption{
696: \label{fig:tmp}
697: For $\gl=0.6$ ($\underline h (0.6)\simeq 0.36$ and $\overline h (0.6)\simeq 0.49$) , the behavior of
698: $\log Z_{2N, \go}$ for $h=0.42$ (A), $0.43$ (C,D) and $0.44$ (B).
699: %The sequence of charges is the same in all the cases.
700: In case A, the polymer is localized with free energy approximately
701: $3\cdot 10^{-6}$:
702: the linear growth is quite clear, but a closer look shows sudden jumps,
703: which correspond to atypically negative stretches of charges.
704: Getting closer to the critical point, case C,
705: the linear growth is still evident, but
706: it is clearly the result of sudden growths followed by slow decays
707: (approximately polynomial with exponent $-1/2$).
708: Case B suggests delocalization: a closer analysis reveals a decay
709: of the type $N ^{-1/2}$, but sharp deviations are clearly visible.
710: %and these deviations
711: %are in reality much larger, since in the graph we have plotted just one point every 10000.
712: Case D is the zoom of the rectangle in the left corner of C.
713: The similarity between B and D
714: make clear that claiming delocalization looking at the behaviour of the partition function is difficult.
715: }
716: \end{figure}
717: 
718: 
719: 
720: \smallskip
721: \subsection{Persistence of the effect of rare stretches}
722: As pointed out in the previous section, there is strong evidence
723: that $h_c(\gl)> \underline{h} (\gl)$.
724: At this stage Fig.~\ref{fig:tmp}  is of particular interest.
725: Notice first of all that in spite of being substantially above $\underline{h}(\cdot)$
726: the copolymer appears to be still localized, see in particular case A.
727: 
728: 
729: 
730: 
731: The rigorous lower bounds that we are able
732: to prove cannot establish localization in the region we are considering.
733: All the same, notice that if one does  not cut the polymer at $T(\go)$,
734: as in the argument above, but at $T(\go)+L$,
735: a  lower bound of the following type
736: \begin{equation}
737: \label{eq:beyondT}
738: Z_{T(\go)+L, \go}\, \stackrel{\text{roughly}}{\ge}\,
739: \text{const.} \frac1{T(\go)^{3/2}}\,
740: \exp\left( -2\gl (q+h)\frac{\log T(\go)}{\Sigma (q)}\right) \, \frac 1 {L^{1/2}},
741: \end{equation}
742: is easily established. Of course we are being imprecise, but we just want to convey
743: the idea, see also Fig.~\ref{fig:ZT}, that after passing through an atypically {\sl negative}
744: stretch of environment ($q>0$), the effect of this stretch decays at most like $L^{-1/2}$,
745: that is the probability that a walk stays  positive for a time $L$.
746: 
747: 
748: At this point we stress that the argument outlined in \S~\ref{sec:lb_outline}
749: and re--used for \eqref{eq:beyondT}
750: may be very well applied to $h> \underline{h}(\gl)$, except that
751: this time it does not suffice for $\eqref{eq:C1}$.
752: But it yields nevertheless  that for $h \in \left(\underline{h}(\gl), \overline{h}(\gl)\right)$
753: the statement $Z_{N,\go}\sim N^{-1/2}$, something a priori expected (for
754: example \cite{cf:BH})
755: in the delocalized regime and true for non disordered systems, is
756: violated. More precisely, one can find a sequence of random
757: times $\{\tau_j\}_j$, $\lim_j \tau _j= \infty$ such that
758: $Z_{\tau_j,\go}\ge {\tau_j}^{-1/2+a}$, $a=a(\gl,h)>0$ (see  Proposition~4.1 in \cite{cf:GT}).
759: These random times are constructed exactly by looking
760: for $q$--atypical stretches as above and one can appreciate
761: such an irregular decay for example in case B of Fig.~\ref{fig:tmp},
762: and this in spite of the fact that the data have been  strongly coarse grained.
763: 
764: \smallskip
765: 
766: Therefore the lower bound
767: \eqref{eq:beyondT}, both in the localized and in the delocalized regime,
768: yields the following  picture:
769: the lower bound we found on  $Z_{N, \go}$ grows suddenly in correspondence of atypical stretches
770: and after that it decays with an exponent $1/2$, up to another atypical stretch.
771: This matches Fig.~\ref{fig:tmp}, at least at a qualitative level, see the caption
772: of the figure.
773: 
774: 
775: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
776: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
777: 
778: 
779: \smallskip
780: \section{The delocalized phase: a path analysis}
781: \label{sec:path}
782: 
783: Let us start with a qualitative observation: if we set the parameters $(\gl,h)$ of the copolymer to $(\gl, h^{(m)}(\gl))$ with $m = 0.9$, then the observed behavior of $\{Z_{N,\go}^{\gl,h}(0)\}_N$ --suitably averaged over blocks in order to eliminate local fluctuations-- is somewhat close to $\text{(const)}/N^{3/2}$. This is true for all the numerically accessible values of $N$ (up to $N\sim 10^8$), at once for a number of values of~$\gl$ and for a great number of typical environments $\go$. Of course this is suggesting that for $m=0.9$ the curve $h^{(m)}(\gl)$ lies in the delocalized region, but it is not easy to convert this qualitative observation into a precise statement, because we do not have a rigorous finite--volume criterion to state that a point $(\gl,h)$ belongs to the delocalized phase (the contrast with the localized phase, see~\eqref{eq:loc_char}, is evident). In other words, we cannot exclude the possibility that the system is still localized but with a characteristic size much larger than the one we are observing.
784: 
785: Nevertheless, the aim of this section is to give an empirical criterion, based on an analysis of the path behavior of the copolymer, that will allow us to provide some more quantitative argument in favor of the fact that the curve $h^{(m)}(\gl)$ lies in the delocalized region even for values of~$m<1$. This of course would entail that the upper bound $\overline h(\gl)$ defined in \eqref{eq:sumup_alt} is not strict.
786: 
787: \smallskip
788: \subsection{Known and expected path behavior}
789: \label{sec:path1}
790: We want to look at the whole \textsl{profile} $\{Z_{N,\go^r}^{\gl,h}(x)\}_{x\in\Z}$ rather than only at $Z_{N,\go^r}^{\gl,h}(0)$, where by $\go^r$ we mean the environment $\go$ in the {\sl
791:  backward direction}, that is $(\go^r)_n := \go_{N+1-n}$ (the reason for this choice is explained in Remark~\ref{rem:backward} below). The link with the path behavior of the copolymer, namely the law of $S_N$ under the polymer measure $\bP_{N,\go^r}^{\gl,h}$, is given by
792: \begin{equation}
793:     \frac{Z_{N,\go^r}^{\gl,h}(x)}{Z_{N,\go^r}^{\gl,h}} = \bP_{N,\go^r}^{\gl,h} (S_N = x)\,.
794: \end{equation}
795: 
796: We have already remarked in \S~\ref{sec:pathwise} of Chapter~\ref{ch:first} that, although the localized and delocalized phases have been defined in terms of free energy, they do correspond to sharply different path behaviors. In the localized phase it is known \cite{cf:Sinai,cf:BisdH} that the laws of $S_N$ under $\bP_{N,\go^r}^{\gl,h}$ are \textsl{tight}, which means that the polymer is essentially at $O(1)$ distance from the $x$--axis. The situation is completely different in the (interior of the) delocalized phase, where one expects that $S_N = O(\sqrt{N})$: in fact the conjectured path behavior (motivated by the analogy with the known results for non disordered models, see in particular \cite{cf:MGO},  \cite{cf:DGZ} and \cite{cf:CGZ}) should be weak convergence under diffusive scaling to the \textsl{Brownian meander process} (that is Brownian motion conditioned to stay positive on the interval $[0,1]$, see \cite{cf:RevYor}). Therefore in the (interior of the) delocalized phase the law of $S_N/\sqrt{N}$ under $\bP_{N,\go^r}^{\gl,h}$ should converge weakly to the corresponding marginal of the Brownian meander, whose law has density $x \exp(-x^2/2) \ind_{(x\ge 0)}$.
797: 
798: In spite of the lack of precise rigorous results,
799: the analysis we are going to describe is carried out under the hypothesis that, in the interior of the delocalized phase, the scaling limit towards Brownian meander holds true (as it will be seen, the numerical results provide a sort of \textsl{a posteriori} confirmation of this hypothesis).
800: 
801: \begin{rem} \label{rem:backward} \rm
802: From a certain point of view attaching the environment backwards does not change too much the model: for example it is easy to check that if one replaces $\go$ by $\go^r$ in \eqref{eq:felim2}, the limit still exists $\bbP(\dd\go)$--a.s. and in $\bbL_1(\dd\bbP)$. Therefore the free energy is the same, because $\{\go^r_n\}_{1\le n \le N}$ has the same law as~$\{\go_n\}_{1\le n \le N}$, for any fixed~$N$.
803: 
804: However, if one focuses  on the law of $S_N$ as a function of~$N$ \textsl{for a fixed environment} $\go$, the behavior reveals to be much smoother under $\bP_{N,\go^r}^{\gl,h}$ than under $\bP_{N,\go}^{\gl,h}$. For instance, under the original polymer measure $\bP_{N,\go}^{\gl,h}$ it is no more true that in the localized region the laws of $S_N$ are tight (it is true only {\sl most of the time}, see \cite{cf:G} for details). The reason for this fact is to be sought in the presence of long {\sl atypical} stretches in every typical~$\go$ (this fact has been somewhat quantified in \cite[Section 4]{cf:GT} and it is at the heart of the approach in Section~\ref{sec:lb}) that are encountered along the copolymer as $N$ becomes larger. Of course the effect of these stretches is very much damped with the backward environment.
805: 
806: A similar and opposite phenomenon takes place also in the delocalized phase. In fancier words, we could say that for fixed $\go$ and as $N$ increases, the way $S_N$ approaches its {\sl limiting behavior} is faster when the environment is attached backwards: it is for this reason that we have
807: chosen to work with $\bP_{N,\go^r}^{\gl,h}$.
808: \end{rem}
809: 
810: \smallskip
811: \subsection{Observed path behavior: a numerical analysis}
812: In view of the above considerations, we choose as a measure of the delocalization of the polymer the
813: $\ell_1$ distance $\bigtriangleup _N^{\gl,h}(\go)$ between the numerically computed profile for a polymer of size $2N$ under $\bP_{2N,\go^r}^{\gl,h}$, and the conjectured asymptotic delocalized profile:
814: \begin{equation}
815:     \bigtriangleup _N^{\gl, h}(\go) := \sum_{x \in 2\Z} \Bigg| \frac{Z_{N,\go^r}^{\gl,h}(x)}{Z_{N,\go^r}^{\gl,h}} \;-\; \frac{1}{\sqrt{2N}} \, \varphi^+ \bigg( \frac{x}{\sqrt{2N}} \bigg) \Bigg|\,,  \qquad \gp^+(x) := x \, e^{-x^2/2} \ind_{(x\ge 0)}\,.
816: \end{equation}
817: Loosely speaking, when the parameters $(\gl,h)$ are in the interior of the the delocalized region we expect $\bigtriangleup _N$ to decrease to~$0$ as $N$ increases, while this certainly will not happen if we are in the localized phase.
818: 
819: \smallskip
820: 
821: The analysis has been carried out at $\gl=0.6$: we recall that the lower and upper bound of \eqref{eq:sumup_alt} give respectively $\underline h(0.6) \simeq 0.36$ and $\overline h(0.6) \simeq 0.49$, while the lower bound we derived with our  test for localization is $h=0.41$, see Table~\ref{tbl:1}. However,
822: as observed in Section \ref{sec:lb}, Fig.~\ref{fig:tmp}, there is numerical evidence that $h=0.43$ is still localized, and for this reason  we have analyzed the values of $h=0.44, 0.45, 0.46, 0.47$ (see below for an analysis on smaller values of~$h$).
823: 
824: 
825: \begin{figure}[h]
826: \begin{center}
827: \leavevmode
828: \epsfysize =7.9 cm
829: \epsfbox{meanders2.eps}
830: \end{center}
831: \caption{
832: \label{fig:F}
833: Graphical representation of the data of Tables~\ref{tbl:F} (on the right) and~\ref{tbl:F2} (on the left). The plotted points are the sample medians against the sample size, the error bars correspond to the confidence intervals given in Tables~\ref{tbl:F} and~\ref{tbl:F2}.
834: }
835: \end{figure}
836: 
837: 
838: For each couple $(\gl, h)$ we have computed $\bigtriangleup _{N}^{\gl,h}(\go)$ for the sizes $N=a\times 10^6$ with $a=1,2,5,10$ and for $500$ independent environments. Of course some type of statistical analysis must be performed on the data in order to decide whether there is a decay of~$\bigtriangleup $ or not. The most direct strategy would be to look at the sample mean of
839: a family of IID variables distributed like $\bigtriangleup _N(\go)$, but it turns out that the fluctuations are too big to get reasonable confidence intervals for this quantity (in other words, the sample variance does not decrease fast enough), at least for the numerically accessible sample sizes. A more careful analysis shows that the variance is essentially due to a \textsl{very small} fraction of data that have \textsl{large} deviations from the mean, while the most of the data mass is quite concentrated.
840: 
841: For this reason we have chosen to focus on the \textsl{sample median} rather than on the sample mean. Table~\ref{tbl:F} contains the results of the analysis (see also Fig.~\ref{fig:F} for a graphical representation): for each value of $h$ we have reported the standard $95\%$ confidence interval for the sample median (see Remark~\ref{rem:conf_int} below for details) for the four different values of~$N$ analyzed. While for $h=0.44$ the situation is not clear, we see that for the values of $h$ greater than $0.45$ there are quantitative evidences for a decrease in~$\bigtriangleup_{N}$: this leads us to the conjecture that the points $(\gl,h)$ with $\gl=0.6$ and $h\ge 0.45$ (equivalently, the points $(\gl,h^{(m)}(\gl))$ with $m \gtrsim 0.876$) lie in the delocalized region.
842: 
843: \begin{table}[h]
844: \begin{center}
845: \begin{tabular}{|c||c|c|c|c|}
846: \hline
847: $h \backslash N(\times 10^6) $ & 1 & 2 & 5 & 10\\
848: \hline
849: \hline
850: 0.44 & [.0603, .0729] & [.0574, .0682] & [.0572, .0689] & [.0570, .0695] \\
851: \hline
852: 0.45 & [.0258, .0286] & [.0207, .0232] & [.0170, .0190] & [.0149, .0171] \\
853: \hline
854: 0.46 & [.0140, .0154] & [.0108, .0116] & [.00792, .00869] & [.00647, .00731] \\
855: \hline
856: 0.47 & [.00905, .00963] & [.00676, .00711] & [.00475, .00508] & [.00364, .00398] \\
857: \hline
858: \end{tabular}
859: \end{center}
860: \bigskip
861: \caption{ \label{tbl:F}
862: The table contains the standard $95\%$ confidence interval for the median of a sample $\{\bigtriangleup _N^{\gl,h}(\go)\}_{\go}$ of size 500, where $\gl=0.6$ and $h,N$ take the different values reported in the table. For the values of $h \ge 0.45$ the decreasing behavior of $\bigtriangleup_{N}$ is quite evident (the confidence intervals do not overlap), see also Fig.~\ref{fig:F}.}
863: \end{table}
864: 
865: As already remarked, these numerical observations cannot rule out the possibility that the system is indeed localized, but the system size is too small to see it. For instance, we have seen that there are evidences for $h=0.43$ to be localized (see case~C of Fig.~\ref{fig:tmp}).
866: In any case, the exponential increasing of $Z_N(0)$ is detectable only at sizes of order$\sim 10^8$, while for smaller system sizes (up to$\sim 10^7$) the qualitative observed behavior of $Z_N(0)$ is rather closer to $(const)/N^{3/2}$, thus apparently suggesting delocalization (see case~D of Fig.~\ref{fig:tmp}).
867: 
868: For this reason it is interesting to look at $\bigtriangleup_N^{0.6,\,h}$ for $h=0.42, 0.43$ and for $N \ll 10^8$. For definiteness we have chosen $N=a\times 10^6$ with $a=1,2,5,10$, performing the computations for $3000$ independent environments: the results are reported in Table~\ref{tbl:F2} (see also Fig.~\ref{fig:F}). As one can see, this time there are clear evidences for an \textsl{increasing} behavior of $\bigtriangleup_N$. On the one hand this fact gives some more confidence on the data of Table~\ref{tbl:F}, on the other hand it suggests that looking at $\{\bigtriangleup_N\}_N$ is a more reliable criterion for detecting (de)localization than looking at $\{Z_N(0)\}_N$.
869: 
870: \begin{table}[h]
871: \begin{center}
872: \begin{tabular}{|c||c|c|c|c|}
873: \hline
874: $h \backslash N(\times 10^5) $ & 1 & 2 & 5 & 10\\
875: \hline
876: \hline
877: 0.42 & [.351, 0.382] & [.480, 0.517] & [.751, 0.794] & [1.01, 1.06] \\
878: \hline
879: 0.43 & [.143, 0.155] & [.165, 0.180] & [.197, 0.215] & [.236, 0.264] \\
880: \hline
881: \end{tabular}
882: \end{center}
883: \bigskip
884: \caption{ \label{tbl:F2}
885: The table contains the standard $95\%$ confidence interval for the median of a sample $\{\bigtriangleup _N^{\gl,h}(\go)\}_{\go}$ of size 3000, where $\gl=0.6$ and $h,N$ take the values reported in the table. For both values of~$h$ an increasing behavior of $\bigtriangleup_{N}$ clearly emerges, see also Fig.~\ref{fig:F} for a graphical representation.}
886: \end{table}
887: 
888: 
889: \begin{rem} \label{rem:conf_int} \rm
890: A confidence interval for the sample median can be obtained in the following general way
891: (the steps below are performed under the assumption that the median
892: is unique, which is, strictly speaking, not true in our case, but it will be clear that
893: a finer analysis would not change the outcome). Let $\{Y_k\}_{1\le k \le n}$ denote a sample of size~$n$, that is the variables $\{Y_k\}_k$ are independent with a common distribution, whose median we denote by~$\xi_{1/2}$:
894: $\bP \left(Y_1 \le \xi_{1/2}\right)=1/2$. Then the variable
895: \begin{equation}
896:     \cN_n := \# \{i \le n:\ Y_i \le \xi_{1/2}\}
897: \end{equation}
898: has a binomial distribution $\cN_n \sim B(n,1/2)$ and when $n$ is large (for us it will be at least~500) we can approximate $\cN_n/n \approx 1/2 + Z/(2\sqrt{n})$, where $Z \sim N(0,1)$ is a standard gaussian. Let us denote the sample quantiles by $\Xi_q$, defined for $q \in (0,1)$ by
899: \begin{equation}
900:     \# \{i \le n:\ Y_i \le \Xi_q\} = \lfloor qn \rfloor\,.
901: \end{equation}
902: If we set $a:= |\Phi^{-1}(0.025)|$ ($\Phi$ being the standard gaussian distribution function) then the random interval
903: \begin{equation}
904:     \Big[\Xi_{\frac{1}{2}-\frac{a}{2\sqrt{n}}},\; \Xi_{\frac{1}{2}+\frac{a}{2\sqrt{n}}}\Big]
905: \end{equation}
906: is a $95\%$ confidence interval for $\xi_{1/2}$, indeed
907: \begin{align}
908:     0.95 &= \bP \big( Z \in [-a,a] \,\big) = \bP \bigg( \frac 12 + \frac{1}{2\sqrt{n}}Z \;\in\; \Big[\frac{1}{2}-\frac{a}{2\sqrt{n}}\;,\; \frac{1}{2}+\frac{a}{2\sqrt{n}} \Big] \bigg) \nonumber \\
909:     &\approx \bP \bigg ( \frac{\cN_n}{n} \in \Big[\frac{1}{2}-\frac{a}{2\sqrt{n}}\;,\; \frac{1}{2}+\frac{a}{2\sqrt{n}} \Big] \bigg) = \bP \bigg( \Xi_{\frac{1}{2}-\frac{a}{2\sqrt{n}}} \le \xi_{1/2} \le \Xi_{\frac{1}{2}+\frac{a}{2\sqrt{n}}} \bigg)\,.
910: \end{align}
911: \end{rem}
912: 
913: 
914: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
915: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
916: 
917: \smallskip
918: \section{An empirical observation on the critical curve}
919: \label{sec:guess}
920: 
921: The key point of this section is that, from a numerical viewpoint,
922: $h_c(\cdot)$ seems very close to $h^{(m)}(\cdot)$, for a suitable
923: value of $m$. Of course any kind of
924:  statement in this direction
925:  requires first of all a procedure to estimate $h_c(\cdot)$ and
926: we explain this first.
927: 
928: Our analysis is based on the following
929: conjecture:
930: \begin{equation}
931: \label{eq:conject_h}
932: (\gl, h)\in \overset{\circ}{\cD} \, \, \Longrightarrow \ \
933: \lim_{N \to \infty }Z_{2N, \go}^{\gl, h} (0) \, =\, 0, \ \bbP \left( \dd \go \right)-\text{a.s.}.
934: \end{equation}
935: The  arguments in Section~\ref{sec:lb}
936: suggest the validity of such a conjecture, which  is  comforted  by the numerical observation.
937: Since, if $(\gl, h)\in \cL$,  $Z_{2N, \go}^{\gl, h} (0)$ diverges (exponentially fast)
938: $\bbP \left( \dd \go \right)$--almost surely and since $Z_{2N, \go}^{\gl, h} (0)$
939: is decreasing  in $h$, we define $\hat h _{N, \go} (\gl)$ as the only $h$ that solves
940: $ Z_{2N, \go}^{\gl, h} (0)=1$.
941: We expect that $\hat h _{N, \go} (\gl)$ converges to $h_c(\gl)$
942: as $N$ tends to infinity, for typical $\go$'s.
943: Of course setting the threshold to the value $1$ is rather arbitrary, but it is
944: somewhat suggested by \eqref{eq:loc_char} and by  the idea behind
945: the proof of \eqref{eq:mainBG} (Proposition  \ref{prop:stopping1} and
946: equation
947: \eqref{eq:T^C}).
948: 
949: \begin{figure}[h]
950: \begin{center}
951: \leavevmode
952: \epsfysize =8.5 cm
953: \psfragscanon
954: \psfrag{x}[c][l]{$\gl$}
955: \psfrag{xg}[c][l]{$\gl$}
956: \psfrag{y}[c][l]{$\hat h _{N,\go}(\gl)$}
957: \psfrag{yg}[c][l]{$\hat h _{N,\go}(\gl)$}
958: \epsfbox{sec5_1.eps}
959: \end{center}
960: \caption{
961: \label{fig:sec5_1}
962: On the left the case of binary symmetric $\go_1$ and on the right
963: the case of $\go_1 \sim N(0,1)$, boths for $N= 3.2\cdot 10^7$.
964: The small circles represent the computed values: the errors on $\hat h _{N,\go}(\gl)$
965: are
966: negligible and the plotted points are at the centers of the circles. The continuous
967: line is instead the curve $h^{(m)}(\cdot)$. In the binary case $m=0.841$ and it has been chosen
968: by solving $h^{(m)}(4)=\hat h _{N,\go}(4)$. In the Gaussian case $m=0.802$, the maximum
969: of $ \hat h _{N,\go}(\gl)/\gl$ for the plotted values of $\gl(>0)$.
970: The rather different values of $\hat m_{N, \go}$ may be somewhat understood
971: both by considering that these two curves have been obtained for a fixed
972: realization of $\go$ and by taking into account the remark at the end of the caption
973: of Table~\ref{tbl:2}: it appears that for  Gaussian charges one needs longer systems in order to get closer
974: to the values of $m$ observed in the binary case (in particular: for the prolongation,
975: with the same random
976: number generator,
977: of the Gaussian $\go$ sample used here up to $N=5\cdot 10^{7}$ one
978: obtains $\hat m_{N, \go}=0.812$).
979: }
980: \end{figure}
981: 
982: 
983: 
984: 
985: 
986: 
987: What we have observed numerically, see Figures~\ref{fig:sec5_1} and \ref{fig:sec5_2},
988: may be summed up by
989: the statement
990: \begin{equation}
991: \label{eq:cnj}
992: \text{there exists } m \text{ such that }
993: \hat h _{N, \go} (\gl) \, \approx h^{(m)}(\gl).
994: \end{equation}
995: Practically this means that $\hat h _{N, \go} (\gl)$, for a set of $\gl$
996: ranging from $0.05$ to $4$, may be fitted with remarkable precision
997: by the one parameter family of functions $\left\{ h^{(m)}(\cdot)\right\}_m$.
998: The fitting value of $m=: \hat m_{N, \go}$ does depend on $N$ and it is essentially increasing.
999: This is of course expected since localization requires a sufficiently large
1000: system (recall in particular Table~\ref{tbl:2} and Fig.~\ref{fig:NCA} -- see the caption of
1001: Fig.~\ref{fig:sec5_1} for the fitting criterion).
1002: We stress that we are presenting results that have been obtained for one
1003: fixed sequence of $\go$: based on what we have observed for example
1004: in Section~\ref{sec:superadd} for different values of $\gl$ one does expect that
1005: for smaller values of $\gl$ one should use larger values of $N$, but
1006:  changing $N$ corresponds to selecting
1007: a longer, or shorter, stretch of $\go$, that is a different sequence of charges
1008: and this may have a rather strong effect on the value of $\hat m_{N, \go}$.
1009: Moreover there is the problem of deciding which $\gl$-dependence to choose. This may explain
1010: the deviations from \eqref{eq:cnj} that are observed for small values of $\gl$, but these
1011: are in any case rather moderate (see Fig.~\ref{fig:sec5_2}).
1012: \smallskip
1013: 
1014:  \begin{figure}[h]
1015: \begin{center}
1016: \leavevmode
1017: \epsfysize =7.1 cm
1018: \psfragscanon
1019: \psfrag{Binary}[c][l]{\small $\go_1 =\pm 1, \, \go_1 \sim -\go_1 $}
1020: \psfrag{Gauss}[c][l]{$\go_1 \, \sim \, N(0,1)$}
1021: \psfrag{x}[c][l]{$\gl$}
1022: \psfrag{xg}[c][l]{$\gl$}
1023: \psfrag{y}[c][l]{$\hat h _{N,\go}(\gl)$}
1024: \psfrag{yg}[c][l]{$\hat h _{N,\go}(\gl)$}
1025: \psfrag{tm1}[c][l]{$r_{N,\go}(\gl)$}
1026: \psfrag{tmp1}[c][l]{$r_{N,\go}(\gl)$}
1027: \epsfbox{sec5_2.eps}
1028: \end{center}
1029: \caption{
1030: \label{fig:sec5_2}
1031: Relative errors $r_{N,\go}(\gl):= \left( h^{(m)}(\gl) -\hat h _{N, \go} (\gl) \right)/\hat h _{N, \go} (\gl)$,
1032: for the value $m=\hat m_{N,\go}$ explained in the caption of Fig.~\ref{fig:sec5_1} and
1033: for the cases of $N=2.5\cdot 10^5$ ($\times$ dots), and $N=3.2\cdot 10^7$ ($+$ dots).
1034: Notice that in the binary case the error is more important for small values of $\gl$ (recall
1035: Table~\ref{tbl:2} and Fig.~\ref{fig:NCA}). Instead for the Gaussian case there is a deviation
1036: both for small and large values of $\gl$: the deviation for large values is due to
1037: the saturation effect explained in the text. Given the fact that $h_{\text{sat}}$, cf. \eqref{eq:sat},
1038: behaves almost surely and to leading order for $N \to \infty$ as $\sqrt{\log N}$ one
1039: understand  why the slow disappearing of the saturation effect has to be expected.
1040: In both graphs the dotted line above the axis is at level $0.01$.
1041: The fitted values for $\hat m_{N, \go}$, $N=  2.5\cdot 10^5$, are $0.821$ in the binary case
1042: and $0.778$ in the Gaussian case.
1043: }
1044: \end{figure}
1045: 
1046: 
1047: 
1048: A source of stronger (and unavoidable) deviations arises in the cases of unbounded charges:
1049: of course if
1050: \begin{equation}
1051: \label{eq:sat}
1052: h\, \ge \,  h_{\text{sat}} \, :=\, \max_{n\in \{1, \ldots, N \}} \left( -(\go_{2n-1}+ \go_{2n})/2 \right),
1053: \end{equation}
1054: then
1055: $Z^{\gl, h}_{2N, \go} (0) <1$, regardless of the value of $\gl$.
1056: Moreover it is immediate to verify that $\lim_{\gl \to \infty } Z^{\gl, h}_{2N, \go}(0) =+\infty$
1057: for $h<  h_{\text{sat}}$ and therefore $\hat h _{N, \go} (\gl)\nearrow h_{\text{sat}} $
1058: as $\gl \nearrow \infty$. We refer to the captions of Fig.~\ref{fig:sec5_2} for
1059: more on this saturation effect.
1060:  \smallskip
1061: 
1062: 
1063: 
1064: 
1065: We have tried also alternative definitions of $\hat h _{N, \go} (\gl)$, namely:
1066: \begin{enumerate}
1067: \item the value of $h$ such that $ Z_{2N, \go}^{\gl, h} =1$ (or a different fixed value);
1068: \item the value of $h$ such that the $\ell_1$ distance between
1069: the distribution of the endpoint and the distribution of the meander, cf. Section~\ref{sec:path}, is smaller than a fixed threshold, for example $0.05$.
1070: \end{enumerate}
1071: 
1072: What we have observed is that \eqref{eq:cnj} still holds.
1073: What is not independent of the criterion is
1074:  $\hat m_{N, \go}$.
1075:  Of course believing deeply in \eqref{eq:cnj} entails
1076:  the expectation  that $\hat m_{N, \go}$ converges to the non random
1077:  quantity $h ^\prime_c (0)$.
1078:   The results reported in this section suggest a value of $h ^\prime_c (0) $ larger than $0.83$
1079:  and the cases presented in Section~\ref{sec:path} suggest that it should be smaller than $0.86$.
1080: 
1081: 
1082: 
1083: 
1084: 
1085: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1086: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1087: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1088: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1089: 
1090: \smallskip
1091: \section{Appendix}
1092: 
1093: \smallskip
1094: \subsection{The algorithm for computing $Z_{N,\go}$}
1095: \label{app:algo}
1096: 
1097: We are going to briefly illustrate the algorithm we used in the numerical computation of the partition function $Z_{N}=Z_{N,\go}^{\gl,h}$. We recall its definition:
1098: \begin{equation} \label{eq:appZ}
1099: Z_N = \bE \Bigg[ \exp \Bigg( -2\gl \sum_{n=1}^N (\go_n + h) \gD_n \Bigg) \Bigg]\,,
1100: \end{equation}
1101: where $\gD_n := (1-\sign(S_n))/2$ and the convention for $\sign(0)$ described in the introduction.
1102: 
1103: Observe that a direct computation of $Z_N$ from (\ref{eq:appZ}) would require to sum the contributions of $2^N$ random walk trajectories, making the problem numerically intractable.
1104:  However, here we can make profitably use of the \textsl{additivity} of our Hamiltonian:
1105:  %(the expression in the argument of the $\exp$ in (\ref{eq:appZ})):
1106:  loosely speaking, if we join together two (finite) random walk segments, the energy of the resulting path is the sum of the energies of the building segments.
1107: 
1108: We can exploit this fact to derive a simple recurrence relation for the sequence of functions $\big\{ \cZ _{M}(y):=Z_{2M}(2y),\ y \in \Z\big\}_{M\in\N}$, where $Z_N(x) = Z_{N,\go}^{\gl,h}(x)$, the latter
1109: defined in \eqref{eq:pinned},
1110: %\begin{equation}
1111: %Z_N(x) = Z_{N,\go}^{\gl,h}(x) = \bE \Bigg[ \exp \Bigg( -2\gl \sum_{n=1}^N (\go_n + h) \gD_n \Bigg);
1112: %S_N=x \Bigg]\,,
1113: %\end{equation}
1114: and we recall that we work  with even values of~$N$. Conditioning on $S_{2M}$ and using the Markov property one easily finds
1115: \begin{equation} \label{eq:apprec}
1116: \cZ_{M+1} (y) = \begin{cases}
1117: \frac14 \cZ_{M}(y+1) \;+\; \frac12 \cZ_{M}(y) \;+\; \frac14 \cZ_{M}(y-1) & y>0 \\
1118: \frac14 \Big[ \cZ_{M}(1) + \cZ_{M}(0) \Big] \;+\; \frac14 \ga_M \Big[ \cZ_{M}(0) + \cZ_{M}(-1) \Big] & y=0 \\
1119: \ga_M \Big[ \frac14 \cZ_{M}(y+1) \;+\; \frac12 \cZ_{M}(y) \;+\; \frac14 \cZ_{M}(y-1) \Big] & y<0
1120: \end{cases} \;,
1121: \end{equation}
1122: where we have put $\ga_M := \exp\big(-2\gl\,(\go_{2M+1} + \go_{2M+2} + 2h)\big)$.
1123: 
1124: From equation (\ref{eq:apprec}) and from the trivial observation that $\cZ_{M}(y)=0$ for $|y|>M$, it follows that $\{\cZ_{M+1}(y),\ y \in \Z\}$ can be obtained from $\{\cZ_{M}(y),\ y \in \Z\}$ with $O(M)$ computations. This means that we can compute $Z_N$  in $O(N^2)$ steps.\footnote{The algorithm just described can be implemented in a standard way: the code we used, written in  C,
1125:  is available on the web page:
1126:  {\tt http://www.proba.jussieu.fr/pageperso/giacomin/C/prog.html}. Graphic representations and standard statistical procedures have been performed
1127: with R \cite{cf:R}. }
1128: 
1129: \medskip
1130: 
1131: We point out that sometimes one is satisfied with \textsl{lower bounds} on $Z_N$, for instance in the statistical text for localization described in Section~\ref{sec:superadd}. In this case the algorithm can be further speeded up by restricting the computation to a suitable set of random walk trajectories. In fact when the system size is~$N$ the polymer is at most at distance $O(\sqrt{N})$ (we recall the discussion in Section~\ref{sec:path} on the path behavior), hence a natural choice to get a lower bound on $Z_N$ is to only take into account the contribution coming from those random walk paths $\{s_n\}_{n\in\N}$ for which
1132: \begin{equation}
1133: -A\sqrt{n} \le s_n \le B \sqrt{n} \qquad \text{for } n \ge N_0\,,
1134: \end{equation}
1135: where $A,B,N_0$ are positive constants. Observe that this is easily implemented in the algorithm described above: it suffices to apply relation (\ref{eq:apprec}) only for $y\in[-A\sqrt M, B\sqrt M]$, while setting $\cZ_{M+1}(y)=0$ for the other values of~$y$. In this way the number of computations needed to obtain $Z_N$ is reduced to~$O(N^{3/2})$.
1136: 
1137: The specific values of $A,B,N_0$ we used in our numerical computations are $3,8,1000$, and we would like to stress that the lower bound on $Z_N$ we got coincides up to the $8^{\text{th}}$ decimal digit with the {\sl true value} obtained applying the complete algorithm.
1138: %This reduced algorithm has permitted us to make computations with systems of size up to $N \sim 10^8$.
1139: 
1140: %A final observation is that, when using the complete algorithm, a major limitation in simulating very big systems is the huge amount of RAM needed. In fact the algorithm needs to operate on the entire vector $\{\cZ_M(y),\ y = -M, \ldots, M\}$, that consequently must be entirely stored.
1141: 
1142: 
1143: \smallskip
1144: A final important remark is that for the results we have reported we have used
1145: the Mersenne--Twister~\cite{cf:MT} pseudo--random number generator.
1146: However we have also tried other pseudo--random number generators
1147: and {\sl true randomness} from {\tt www.random.org}:
1148: the results appear not to depend on the generator.
1149: 
1150: 
1151: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1152: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1153: 
1154: \smallskip
1155: \subsection{Proof of the lower bound on $h_c$}
1156: \label{app:prooflb}
1157: We are going to give a detailed proof of the lower bound \eqref{eq:mainBG} on the critical curve, together with some related result. We stress that this appendix can be made substantially lighter
1158: if one is interested only in the {\sl if} part of Proposition~\ref{prop:stopping1}. In this case
1159: the first part of this appendix is already contained in the first part of \S~\ref{sec:lb_outline}, up
1160: to \eqref{eq:forappB}, and it suffices to look at the proof of the lower bound
1161: starting from page~\pageref{sec:appB2}.
1162: 
1163: 
1164: We recall that $Z_{N,\go}^{\gl,h}(0)$ is the partition function corresponding to the polymer pinned at its right endpoint, see \eqref{eq:pinned}, and $T^C=T^{C}(\go)$ is the first $N$ for which $Z_{N,\go}(0)\ge C$, see \eqref{eq:T^C}. In particular, for all $\go$ such that $T^C(\go) < \infty$ we have
1165: \begin{equation} \label{eq:stopping_maj}
1166: Z_{T^C(\go),\go}^{\gl,h}(0) \ge C\,.
1167: \end{equation}
1168: We will also denote by $\cF_n := \gs(\go_1,\ldots,\go_n)$ the natural filtration of the sequence $\{\go_n\}_{n\in\N}$.
1169: 
1170: \medskip
1171: 
1172: \subsubsection{A different look at (de)localization}
1173: We want to show that (de)localization can be read from $T^C$. We introduce some notation: given an increasing, $2\N$--valued sequence $\{t_i\}_{i\in\N}$, we set $t_0:=0$ and $\zeta_N:=\max\{k: t_k \le N\}$. Then we  define
1174: \begin{align} \label{eq:low_b}
1175: \begin{split}
1176: \widehat{Z}_{N,\go}(0) = \widehat{Z}_{N,\go}^{\{t_i\},\gl,h}(0) & \;:=\; \bE \left[ e^{-2 \gl \sum_{n=1}^N \left( \go_n +h\right) \Delta_n}; \, S_{t_1}=0,\, \ldots,\, S_{t_{\zeta_N}}=0,\, S_N= 0 \right] \\
1177: &\;=\; \prod_{i=0}^{\zeta_N-1} Z_{t_{i+1}-t_i,\theta^{t_i}\go}^{\gl,h}(0) \,\cdot\, Z_{N-t_{\zeta_N}(\go),\theta^{t_{\zeta_N}}\go}^{\gl,h}(0)\,,
1178: \end{split}
1179: \end{align}
1180: and we recall that $\theta$ denotes the translation on the environment.
1181: One sees immediately that $\widehat{Z}_{N,\go}(0)\le Z_{N,\go}(0)$.
1182: We first establish a preliminary result.
1183: 
1184: \begin{lemma} \label{lem:chop}
1185: If the sequence $\{t_i\}_i$ is such that $\zeta_N/N \to 0$ as $N\to\infty$, then
1186: \begin{equation}
1187: \lim_{N\to\infty} \frac 1N \log \widehat{Z}_{N,\go}^{\{t_i\},\gl,h}(0) \;=\; \tf(\gl,h)\,,
1188: \end{equation}
1189: both $\bbP(\dd\go)$--a.s. and in $\bbL_1(\bbP)$.
1190: \end{lemma}
1191: 
1192: \proof
1193: By definition we have $Z_{N,\go}(0) \ge \widehat{Z}_{N,\go}(0)$. On the other hand, we are going to show that
1194: \begin{equation} \label{eq:rough_maj}
1195: Z_{N,\go}^{\gl,h}(0) \;\le\; 4^{\zeta_N} \, A^{2\zeta_N} \, \left( \prod_{i=1}^{\zeta_N} (t_i-t_{i-1}) \cdot (N-t_{\zeta_N}) \right)^{3} \,  \widehat{Z}_{N,\go}^{\{t_i\},\gl,h}(0)\,,
1196: \end{equation}
1197: where $A$ is a positive constant. To derive this bound, we resort to the equation \eqref{eq:reducttoexc} that expresses $Z_{N,\go}(0)$ in terms of random walk excursions. We recall that $K(2n)$ is the discrete probability density of the first return time of the walk $S$ to $0$, and that $K(t) \ge 1/(A\,t^{3/2})$, $t\in 2\N$, for some positive constant~$A$: it follows that for $a_1, \ldots, a_k \in 2\N$
1198: \begin{equation} \label{eq:app_bound}
1199: K(a_1 + \ldots + a_k) \; \le \; 1 \; \le \; A^k \, (a_1 \cdot \ldots \cdot a_k)^{3/2} \, K(a_1) \cdot \ldots \cdot K(a_k)\,.
1200: \end{equation}
1201: This gives us an upper bound to the entropic cost needed to split a random walk excursion of length $(a_1 + \ldots + a_k)$ into $k$~excursions of lengths $a_1, \ldots, a_k$.
1202: 
1203: Now let us come back to the second line of \eqref{eq:reducttoexc}, that can be rewritten as
1204: \begin{equation}
1205: \label{eq:app_sum}
1206: Z_{N,\go}(0) \,=\, \sum_{\{x_i\} \subseteq \{0, \ldots, N\} \cap 2\N} G(\{x_i\})\,.
1207: \end{equation}
1208: A first observation is that if we restrict the above sum  to the $\{x_i\}$ such that
1209: $\{x_i\} \supseteq \{t_i\}$, then we  get  $\widehat{Z}_{N,\go}^{\{t_i\}}(0)$. Now for each $\{x_i\}$ we
1210: aim at finding an  upper bound on the term $G(\{x_i\})$ of the form
1211:  $c \cdot G(\{x_i\} \cup \{t_i\})$ for some $c>0$ not depending on $\{x_i\}$.
1212:  Each term $G(\{x_i\})$, see \eqref{eq:reducttoexc}, is the product of two terms: an entropic part depending on~$K(\cdot)$ and an energetic part depending on~$\varphi(\cdot)$. Replacing the entropic part costs no more than
1213: \begin{equation}
1214:     c_{\text{ent}} \, :=\,  A^{2\zeta_N} \, \left( \prod_{i=1}^{\zeta_N} (t_i-t_{i-1}) \cdot (N-t_{\zeta_N}) \right)^{3}\,,
1215: \end{equation}
1216: thanks to~\eqref{eq:app_bound}. On the other hand, the cost for replacing the energetic part is easily
1217: bounded above by
1218: \begin{equation}
1219:     c_{\text{energy}} \, :=\, 2^{\zeta_N}\,,
1220: \end{equation}
1221: so that the bound $G(\{x_i\}) \le c \cdot G(\{x_i\} \cup \{t_i\})$ holds true with $c:= c_{\text{ent}} \, c_{\text{energy}}$. Replacing in this way each term in the sum in the r.h.s. of \eqref{eq:app_sum}, we are left with a sum of terms $G(\{y_i\})$ corresponding to sets $\{y_i\}$ such that $\{y_i\} \supseteq \{t_i\}$. It remains to count the {\sl multiplicity} of any such~$\{y_i\}$, that is how many original sets $\{x_i\}$ are such that $\{x_i\} \cup \{t_i\} = \{y_i\}$. Sets $\{x_{i}\}$ satisfying this last condition must differ only for a subset of $\{t_{i}\}$, hence the sought multiplicity is $2^{\zeta_N}$ (the cardinality of the parts of $\{t_{i}\}$) and the bound \eqref{eq:rough_maj} follows.
1222: 
1223: Therefore we get
1224: \begin{align}
1225: \bigg| \frac{\log \widehat{Z}_{N,\go}^{\{t_i\},\gl,h}(0)}{N}  -  \frac{\log Z_{N,\go}^{\gl,h}(0)}{N} \bigg| & \;\le\; (2\log 2 A) \frac{\zeta_N}{N} \;+\; 3\, \frac{1}{N} \,\log \left( \prod_{i=1}^{\zeta_N} (t_i-t_{i-1}) \cdot (N-t_{\zeta_N}) \right) \\
1226: &\;\le\; (2\log 2A) \frac{\zeta_N}{N} \;+\; 3 \, \frac{\zeta_N+1}{N} \, \log \bigg(\frac{N}{\zeta_N+1}\bigg)\,,
1227: \nonumber
1228: \end{align}
1229: where in the second inequality we have made use of the elementary fact that once the sum of $k$ positive numbers is fixed, their product is maximal when all the numbers coincide (for us $k=\zeta_N+1$). Since by hypothesis $\zeta_N/N \to 0$ as $N\to\infty$, the Lemma is proved.\qed
1230: 
1231: \bigskip
1232: 
1233: Now we are ready to prove the characterization of $\cL$ and $\cD$ in terms of $T^C$.
1234: Fix any $C>1$.
1235: 
1236: \smallskip
1237: 
1238: \begin{proposition} \label{prop:stopping1}
1239: A point $(\gl,h)$ is localized, that is $h < h_c(\gl)$, if and only if $\bbE[T^C]<\infty$.
1240: \end{proposition}
1241: 
1242: \proof
1243: We set $\cA:=\{\go: T^C(\go)<\infty\}$. Observe that for $\go\in \cA^{\complement}$ we have $Z_{N,\go}(0) \le C$ for every $N\in2\N$, and consequently $\log Z_{N,\go}^{\gl,h}(0)/N \to 0$ as $N\to\infty$.
1244: 
1245: Consider first the case when the random variable $T^C$ is defective, that is $\bbP[\cA^{\complement}]>0$ (this is a particular case of $\bbE[T^C]=\infty$). Since we know that $\log Z_{N,\go}^{\gl,h}(0)/N \to \tf(\gl,h)$, $\bbP(\dd\go)$--a.s., from the preceding observation it follows that $\tf(\gl,h)=0$ and the Proposition is proved in this case.
1246: 
1247: \smallskip
1248: 
1249: Therefore in the following we can assume that $T^C$ is proper, that is $\bbP(\cA)=1$, so that equation (\ref{eq:stopping_maj}) holds for almost every~$\go$. Setting $\theta^{-1}\cA := \{\go : \theta\go \in \cA\}$, we have $\bbP\left(\theta^{-1}\cA\right)=1$ since $\bbP$ is $\theta$--invariant, and consequently $\bbP\left(\cap_{k=0}^\infty \theta^{-k}\cA
1250:  \right)=1$, which amounts to saying that (\ref{eq:stopping_maj}) can be actually strengthened to
1251: \begin{equation} \label{eq:im_stopping_maj}
1252: Z_{T^C(\theta^k\go),\theta^k\go}^{\gl,h}(0) \ge C \qquad \forall k \ge 0,\ \bbP(\dd\go)\text{--a.s.}\,.
1253: \end{equation}
1254: Observe that the sequence $\{( \theta^{T^C(\go)} \go )_n\}_{n\in\N}$ has the same law as $\{\go_n\}_{n\in\N}$ and it is independent of $\cF_{T^C}$. We can define inductively an increasing sequence of stopping times $\{T_n\}_{n\in\N}$ by setting $T_0:=0$ and $T_{k+1}(\go) - T_k(\go) := T^C(\theta^{T_k(\go)}\go) =: S_k(\go)$. We also set $\zeta_N(\go) := \max \{n:\ T_n(\go) \le N\}$. Since $\{S_k\}_{k\in\N}$ is an IID sequence, by the strong law of large numbers we have that, $\bbP(\dd\go)$--a.s., $T_n(\go)/n \to \bbE[T^C]$ as $n\to\infty$, and consequently $\zeta_N(\go)/N \to 1/\bbE[T^C]$ as $N\to\infty$ (with the convention that $1/\infty=0$).
1255: 
1256: Now let us consider the lower bound $\widehat{Z}_{N,\go}(0)$ corresponding to the sequence $\{t_i\}=\{T_i(\go)\}$: from \eqref{eq:low_b} and \eqref{eq:im_stopping_maj} we get that $\bbP(\dd\go)$--a.s.
1257: \begin{align} \label{eq:major}
1258: \begin{split}
1259: \widehat{Z}_{N,\go}^{\{T_i(\go)\},\gl,h}(0) &\;=\; \prod_{i=0}^{\zeta_N(\go)-1} Z_{T^C(\theta^{T_i}\go),\theta^{T_i}\go}^{\gl,h}(0) \,\cdot\, Z_{N-T_{\zeta_N(\go)}(\go),\theta^{T_{\zeta_N(\go)}}\go}^{\gl,h}(0) \\
1260: &\;\ge\; C^{\zeta_N(\go)} \cdot \frac{c}{N^{3/2}}\,,
1261: \end{split}
1262: \end{align}
1263: where $c$ is a positive constant (to estimate the last term we have used the  lower bound $Z_k(0) \ge c/k^{3/2}$, cf. \eqref{eq:step_deloc}), and consequently
1264: \begin{equation}
1265: \tf(\gl,h) \;=\; \lim_{N\to\infty} \frac{\log Z_{N,\go}^{\gl,h}(0)}{N} \;\ge\; \liminf_{N\to\infty} \frac{\log \widehat{Z}_{N,\go}^{\{T_i(\go)\},\gl,h}(0)}{N} \;\ge\; \frac{\log C}{\bbE[T^C]}\,.
1266: \end{equation}
1267: It follows that if $\bbE[T^C]<\infty$  then $\tf(\gl,h)>0$, that is $(\gl,h)$ is localized.
1268: 
1269: \smallskip
1270: 
1271: It remains to consider the case $\bbE[T^C]=\infty$, and we want to show that this time $\widehat{Z}_{N,\go}(0)$, defined in \eqref{eq:major}, gives a null free energy. In fact, as $T^C(\eta)$ is defined as the \textsl{first} $N$ such that $Z_{N,\eta}(0) \ge C$, it follows that $Z_{T^C(\eta),\eta}(0)$ cannot be much greater than $C$. More precisely, one has that
1272: \begin{equation}
1273: Z_{T^C(\eta),\eta}(0) \le  C \,\exp(2\gl|\eta_{T^C(\eta)-1} + \eta_{T^C(\eta)}|)\,,
1274: \end{equation}
1275: and from the first line of \eqref{eq:major} it follows that
1276: \begin{equation}
1277: \frac 1N \log \widehat{Z}_{N,\go}(0) \;\le\; \frac{\zeta_N(\go)+1}{N} \log C \;+\;  \frac{2\gl}{N} \sum_{i=1}^{\zeta_N(\go)} \Big( |\go_{T_i(\go)}| + |\go_{T_i(\go)-1}| \Big)\,.
1278: \end{equation}
1279: We estimate the second term in the r.h.s. in the following way:
1280: \begin{align}
1281:     & \frac 1N \sum_{i=1}^{\zeta_N(\go)} \Big( |\go_{T_i(\go)}| + |\go_{T_i(\go)-1}| \Big) = \frac 1N \sum_{k=1}^{N} \ind_{\{\exists i:\, T_i(\go) = k\}} \Big( |\go_k| + |\go_{k-1}| \Big)\nonumber  \\
1282:     & \qquad \le \left( \frac 1N \sum_{k=1}^{N} \ind_{\{\exists i:\, T_i(\go) = k\}} \right) ^{1/2} \left( \frac 1N \sum_{k=1}^{N} \Big( |\go_k| + |\go_{k-1}| \Big)^2 \right)^{1/2}\\
1283:     &\qquad \le \sqrt{\frac{\zeta_N(\go)}{N}} \cdot 2 \sqrt{ \frac 1N \sum_{k=1}^{N} |\go_k|^2} \le A \sqrt{\frac{\zeta_N(\go)}{N}} \,, \nonumber
1284: \end{align}
1285: for some positive constant~$A=A(\go)$ and  eventually as $N \to \infty$, having used the Cauchy--Schwartz inequality and the law of large numbers for the sequence $\{|\go_k|^2\}_{k\in\N}$. Therefore
1286: \begin{equation}
1287:     \frac 1N \log \widehat{Z}_{N,\go}(0) \;\le\; \frac{\zeta_N(\go)+1}{N} \log C \;+\; 4\gl A \sqrt{\frac{\zeta_N(\go)}{N}}\,,
1288: \end{equation}
1289: and since $\bbE[T^C]=\infty$ implies $\zeta_N(\go)/N \to 0$, $\bbP(\dd\go)$--a.s., we have $\log \widehat{Z}_{N,\go}(0) /N \to 0$, $\bbP(\dd\go)$--a.s.. Then Lemma~\ref{lem:chop} allows us to conclude that $\tf (\gl,h)=0$, and the proof of the Proposition is completed.\qed
1290: 
1291: 
1292: %since by hypothesis $\bbE[\exp(|\go_1|)]<\infty$, it follows that there exists a $\bbP(\dd\go)$--a.s. finite random constant $\cC(\go)$ such that $|\go_n| \le \cC(\go) \log n$, for every $n\in\N$. This observation yields
1293: %\begin{align}
1294: %\frac 1N \sum_{i=1}^{\zeta_N(\go)} |\go_{T_i(\go)}| \le \frac{\cC(\go)}{N} \prod_{i=1}^{\zeta_N(\go)} \log T_i(\go) \cdot ...
1295: %\end{align}
1296: 
1297: \medskip
1298: 
1299: \subsubsection{Proof of the lower bound on $h_c$}
1300: \label{sec:appB2}
1301: To prove equation \eqref{eq:mainBG}, we are going to build, for every $(\gl,h)$ such that $h < \underline h(\gl)$, a random time $T$ such that $\bbE[T]<\infty$ and $Z_{T(\go),\go}^{\gl,h}(0) \ge C$, for some $C>1$. It follows that $T^C \le T$, yielding that $\bbE[T^C]<\infty$ and by Proposition~\ref{prop:stopping1} $(\gl,h)$ is localized, that is, $\underline h(\gl) \le h_c(\gl)$.
1302: 
1303: \smallskip
1304: 
1305: Given $M\in 2\N$
1306: and $q<-h$, we start defining the stopping time
1307: \begin{equation}
1308: \tau_M(\go) = \tau_{M,q}(\go) := \inf \bigg\{n\in2\N:\ \exists k \in 2\N,\ k \ge M:\ \frac{\sum_{i=n-k+1}^n \go_i}{k} \le q \bigg\}\,.
1309: \end{equation}
1310: This is the first instant at which a $q$--atypical stretch of length at least $M$ appears along the sequence $\go$. The asymptotic behavior of $\tau_M$ is given by Theorem 3.2.1 in \cite[\S~3.2]{cf:DZ} which says that $\bbP(\dd\go)$--a.s.
1311: \begin{equation} \label{eq:as_tau}
1312: \frac{\log \tau_{M}(\go)}{M} \to \gS(q) \qquad \text{as } M\to\infty\,,
1313: \end{equation}
1314: where $\gS(q)$ is Cramer's Large Deviations functional for $\go$, \eqref{eq:cramer}.
1315: We also give a name to the shortest of the terminal stretches in the definition of $\tau_M$:
1316: \begin{equation}
1317: R_{M}(\go) = R_{M,q}(\go) := \inf \bigg\{k \in 2\N,\ k \ge M:\ \frac{\sum_{i=\tau_M-k+1}^{\tau_M} \go_i}{k} \le q \bigg\}\,,
1318: \end{equation}
1319: and it is not difficult to realize that $R_M \le 2M$.
1320: 
1321: 
1322: \smallskip
1323: 
1324: We are ready to give a simple lower bound on the partition function of size $\tau_{M,q}$ (for any $M\in 2\N$ and $q<-h$): it suffices to consider the contribution of the trajectories that are negative in correspondence of the last (favorable) stretch of size $R_M$, and stay positive the rest of the time. Recalling that
1325: we use $K(\cdot)$ for the discrete density of the first return time to the origin and that by
1326: \eqref{eq:asympt} we have $K(2n)\ge c/n^{-3/2}$ for a constant $c>0$,
1327: we estimate
1328: \begin{equation}\label{eq:needs}
1329: \begin{split}
1330: Z_{\tau_M(\go), \go}^{\gl,h}(0) & \ge \frac 1 4 \, K\left({\tau_M - R_M}\right) \, K\left(
1331: {R_M}\right) \, e^{-2\gl (q+h) R_M} \ge \frac{c^2}{4 \tau_M^{3/2} (2M)^{3/2}} e^{-2\gl (q+h) M} \\
1332: & \ge c'\, \exp \bigg\{ \frac 3 2 M \bigg[ (-4\gl /3)q - \frac{\log \tau_M}{M} - ( 4 \gl/3) h - \frac{\log M}{M} \bigg] \bigg\} \,,
1333: \end{split}
1334: \end{equation}
1335: where $c' := c^2/(8\sqrt{2})$.
1336: 
1337: Having in mind (\ref{eq:as_tau}), we define a random index $\ell=\ell_{A,\gep,q}$ depending on the two parameters $A \in 2\N,\ \gep>0$ and on $q$:
1338: \begin{equation} \label{eq:def_ell}
1339: \ell(\go)  = \ell_{A,\gep,q}(\go) := \inf \bigg\{ k\in 2 \N,\ k \ge A:\ \frac{\log \tau_{k,q}(\go)}{k} \le \gS(q) + \gep \bigg\} \,,
1340: \end{equation}
1341: and we finally set
1342: \begin{equation}\label{eq:def_T_cgg}
1343:     T (\go) = T_{A,\gep,q} (\go) := \tau_{\ell(\go)}(\go)\,.
1344: \end{equation}
1345: Then for the partition function of size $T(\go)$ we get
1346: \begin{equation} \label{eq:almost_maj}
1347: Z_{T(\go), \go}^{\gl,h}(0) \ge c'\, \exp \bigg\{ \frac 3 2 A \bigg[ (-4\gl /3)q - \gS(q) - (4\gl /3)h - \frac{\log A}{A} - \gep \bigg] \bigg\} \,.
1348: \end{equation}
1349: 
1350: The fact that $\bbE[T_{A,\gep,q}]<\infty$ for any choice of $A,\gep,q$ (with $q<-h$) is proved in Lemma~\ref{lem:stopping2} below. It only remains to show that for every fixed $(\gl,h)$ such that $h<\underline h(\gl)$, or equivalently
1351: \begin{equation}
1352: \label{eq:cond_low}
1353: (4\gl/3)h < \log \M(-4\gl/3)\,,
1354: \end{equation}
1355: the parameters $A,\gep,q$ can be chosen such that the right--hand side
1356:  of equation (\ref{eq:almost_maj}) is greater than~$1$.
1357: 
1358: \smallskip
1359: 
1360: The key point is the choice of  $q$. Note that
1361: the generating function $ \M(\cdot)$ is smooth, since  finite on the whole real line.
1362:  Moreover for all $ \gl \in\R$ there exists some $q_0\in\R$ such that
1363: \begin{equation}
1364: \log \M(-4\gl/3) = (-4\gl/3) q_0 - \gS(q_0)\,,
1365: \end{equation}
1366: and from (\ref{eq:cond_low}) it follows that $q_0 < -h$. Therefore we can take $q=q_0$, and equation (\ref{eq:almost_maj}) becomes
1367: \begin{equation}
1368: \label{eq:lbonZTom}
1369: Z_{T(\go), \go}^{\gl,h}(0) \ge c'\, \exp \bigg\{ \frac 3 2 A \bigg[ \log \M(-4\gl/3) - (4\gl/3) h - \frac{\log A}{A} - \gep \bigg] \bigg\} \,.
1370: \end{equation}
1371: It is now clear that for every $(\gl,h)$, such that (\ref{eq:cond_low}) holds, by choosing
1372: $\gep$ sufficiently small and $A$ sufficiently large, the right--hand side of \eqref{eq:lbonZTom} is greater than~1, and the proof of \eqref{eq:mainBG} is complete.
1373: 
1374: \medskip
1375: 
1376: \begin{lemma}
1377: \label{lem:stopping2}
1378: For every $A\in 2\N$,  $\gep > 0$ and $ q <-h$
1379: the random variable $T(\go)=T_{A,\gep,q}(\go)$ defined
1380: below \eqref{eq:def_ell} is integrable: $\bbE[T]<\infty$.
1381: \end{lemma}
1382: 
1383: \medskip
1384: 
1385: \proof
1386: By the definition (\ref{eq:def_ell}) of $\ell=\ell_{A,\gep,q}$ we have
1387: \begin{equation}
1388: T_{A,\gep,q} \le \exp\big( (\gS(q) + \gep)\, \ell_{A,\gep,q} \big)\,,
1389: \end{equation}
1390: so it suffices to show that for any $\gb>0$ the random variable $\exp(\gb\, \ell_{A,\gep,q})$ is integrable.
1391: 
1392: For any $l\in2\N$, we introduce the IID sequence of random variables $\{Y_n^{l}\}_{n\in\N}$ defined by
1393: \begin{equation}
1394: Y^{l}_n := \frac 1 {l}{\sum_{i=(n-1)l+1}^{n l} \go_i}\,.
1395: \end{equation}
1396: By Cramer's Theorem \cite{cf:DZ} we have that for any fixed $q<0$ and $ \gep>0$ there exists  $l_0$ such that $\bbP\left(Y_1^{l} \le q\right) \ge e^{-l (\gS(q) + \gep/2)}$ for every $l\ge l_0$.
1397: By (\ref{eq:def_ell}) have that
1398: \begin{equation}
1399: \begin{split}
1400: \{\ell > l\} \subseteq \left\{\tau_l > \exp((\gS(q)+\gep) l)\right\}
1401: \subseteq \inter_{i=1}^{\lfloor M/l \rfloor} \{Y_i^{l} > q\}\,,
1402: \end{split}\end{equation}
1403: with $M:=\exp((\gS(q)+\gep) l)$, so that
1404: \begin{equation}
1405: \begin{split}
1406: \bbP\left(\ell > l\right)
1407:   \le \left( 1- e^{-l (\gS(q) + \gep/2)} \right)^{\lfloor M/l \rfloor}
1408:   &\le
1409:   \exp \left( - \lfloor M/l \rfloor e^{-l (\gS(q) + \gep/2)} \right)
1410:   \\
1411: & \le  \exp \left(- \exp\left(l \gep/4\right)\right),
1412: \end{split}
1413: \end{equation}
1414: where the last step holds if $l$ is sufficiently large   (we have also used
1415:  $1-x \le e^{-x}$). Therefore
1416: \begin{equation}
1417: \bbP\left(\exp(\gb\,\ell) > N\right) = \bbP\left(\ell > (\log N) /\gb \right)
1418: \le  \exp \left( - N ^{\gep / 4\gb }\right),
1419: \end{equation}
1420: when $N$ is large, and the proof is complete. \qed
1421: 
1422: 
1423: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1424: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1425: 
1426: 
1427: 
1428: