math0611242/ht.tex
1: %Created:               Wed 02 Aug 2006 10:35:58 CEST
2: %Last Modified:         Tue 07 Nov 2006 12:02:16 CET
3: %<<< Preamble
4: \documentclass[a4paper,12pt,oneside,reqno]{amsart}
5: \usepackage[headinclude,DIV13]{typearea}
6: \areaset{15.5cm}{25cm}
7: %\usepackage{a4wide}
8: %\textheight 48\baselineskip
9: \usepackage{amsfonts,amssymb,amsmath,amsthm}
10: %\usepackage[color,notcite]{showkeys}
11: %\usepackage[utf8]{inputenc}
12: 
13: \overfullrule 1mm
14: 
15: \newtheorem{theorem}{Theorem}[section]
16: \newtheorem{lemma}[theorem]{Lemma}
17: \newtheorem{proposition}[theorem]{Proposition}
18: \newtheorem{corollary}[theorem]{Corollary}
19: \newtheorem{exercise}{Exercise}
20: 
21: \theoremstyle{definition}
22: \newtheorem{definition}[theorem]{Definition}
23: \newtheorem{assumption}[theorem]{Assumption}
24: \newtheorem{condition}{Condition}
25: \newtheorem{solution}{Solution}
26: \newtheorem{question}{Question}
27: 
28: \theoremstyle{remark}
29: \newtheorem*{remark}{Remark}
30: \newtheorem*{hint}{Hints}
31: 
32: \numberwithin{equation}{section}
33: 
34: %\usepackage{verbatim}
35: %\let\solution=\comment
36: %\let\endsolution=\endcomment
37: 
38: \def\todo#1{\marginpar{\raggedright\footnotesize #1}}
39: \def\change#1{{\color{green}\todo{change}#1}}
40: 
41: \def\vagconv{\overset v\to }
42: \def\ppconv{\overset {pp} \to }
43: \def\weakconv{\overset w\to }
44: 
45: \def\laweq{\overset{\mathrm{law}}=}
46: 
47: \def\Max{{\mathrm{max}}}
48: \def\Min{{\mathrm{min}}}
49: \def\supp{\mathop{\mathrm{supp}}\nolimits}
50: \def\Id{\mathrm{Id}}
51: \def\dist{\mathop{\rm dist}\nolimits}
52: \def\supp{\mathop{\rm supp}\nolimits}
53: \def\sign{\mathop{\rm sign}\nolimits}
54: \def\psum{\mathop{\sum\nolimits'}}
55: \def\d{\mathrm{d}}
56: \def\dd{\mathtt{d}}
57: \def\bbone{{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l}
58:           {\rm 1\mskip-4.5mu l} {\rm 1\mskip-5mu l}}}
59: %\def\bbone{\chi }
60: 
61: \def\<{\langle}
62: \def\>{\rangle}
63: 
64: %>>>
65: 
66: 
67: \begin{document}
68: \title{Hitting time of large subsets of the hypercube}
69: 
70: \author[J. \v Cern\'y]{Ji\v r\'\i~\v Cern\'y}
71: \address{Ji\v r\'\i~\v Cern\'y\\
72:   \'Ecole Polytechnique F\'ed\'erale de Lausanne\\
73:   1015 Lausanne\\
74:   Switzerland
75:   }
76: \email{jiri.cerny@epfl.ch}
77: 
78: \author[V. Gayrard]{V\'eronique Gayrard}
79: \address{V\'eronique Gayrard\\Laboratoire d'Analyse, Topologie, Probabilit\'es\\
80: CMI, 39 rue Joliot-Curie\\
81: 13453 Marseille Cedex}
82: \email{gayrard@latp.univ-mrs.fr, veronique@gayrard.net}
83: 
84: \subjclass[2000]{60J10, 60K37, 82D30}
85: \keywords{Hitting time, random walk, hypercube, aging}
86: 
87: \date{\today}
88: \begin{abstract}
89:   We study the simple random walk on the $n$-dimensional hypercube, in
90:   particular its hitting times of large (possibly random) sets. We give
91:   simple conditions on these sets ensuring that the properly-rescaled
92:   hitting time is asymptotically exponentially distributed, uniformly in
93:   the starting position of the walk. These conditions are then verified for
94:   percolation clouds with densities that are much smaller than
95:   $(n \log n)^{-1}$. A main motivation behind this paper is the study of
96:   the so-called aging phenomenon in the Random Energy Model (REM),
97:   the simplest model of a mean-field spin glass. Our results allow
98:   us to prove aging in the REM for all temperatures,
99:   thereby extending earlier results to their optimal temperature domain.
100: 
101: \end{abstract}
102: 
103: \maketitle
104: 
105: 
106: %%%% START OF ht1.tex %%%%
107: \section{Introduction}
108: \label{s:introduction}
109: Let $\mathcal V_n$ be the $n$-dimensional hypercube,
110: $\mathcal V_n=\{0,1\}^n$. We equip $\mathcal V_n$ with
111: the metric
112: \begin{equation}
113:   d(x,y)=\sum_{i=1}^n \bbone\{x(i)\neq y(i)\},
114: \end{equation}
115: where $x(i)$ are the coordinates of $x\in \mathcal V_n$. Let further $Y_n$ be
116: the simple random walk on $\mathcal V_n$. That is,  $Y_n$ is the
117: discrete-time Markov chain with state space $\mathcal V_n$ whose
118: transition probabilities are given by
119: $\mathbb P[Y_n(k+1)=y|Y_n(k)=x]=n^{-1}$ if $d(x,y)=1$, and $0$ otherwise. We
120: write $\mathbb P_x$ for the distribution of $Y_n$ conditioned on $Y_n(0)=x$.
121: For $A\subset \mathcal V_n$ we define the hitting time of $A$ by
122: \begin{equation}
123:   H_n(A)=\min\{k\ge 0: Y_n(k)\in A\}.
124: \end{equation}
125: 
126: We are interested in the distribution of the hitting time of large random
127: subsets of the hypercube. Specifically, let $\rho \in [0,1] $. We say that
128: the set $A\subset \mathcal V_n$ is a \textit{percolation cloud} on
129: $\mathcal V_n$ with density $\rho $ if  each site $x\in \mathcal V_n$ is
130: in $A$ with probability $\rho $ independently of all others.
131: 
132: 
133: Our main aim is to prove the following theorem.
134: \begin{theorem}
135:   \label{t:perc}
136:   Let $\bar m(n)$ be such that
137:   \begin{equation}
138:     \label{e:condm}
139:     n \log n \ll \bar m(n) \ll 2^{n}(\log n)^{-1}
140:   \end{equation}
141:   and let $A_n$ be a sequence of percolation clouds on $\mathcal V_n$ with
142:   densities $\bar m(n)^{-1}$ defined on a common probability space
143:   $(\Omega,F, P)$.
144:   Then, for all $a>0$,
145:   \begin{equation}
146:     \label{e:thmjedna}
147:     \lim_{n\to\infty}
148:     \max_{x\in \mathcal V_n}
149:     \Big |
150:     \mathbb P_x \big[H_n(A_n\setminus x)\ge a \bar m(n)\big]-e^{-a}
151:     \Big |
152:     =0\,,\quad P\text{-a.s.}
153:   \end{equation}
154:   In words, the distribution of the normalised hitting time of
155:   $A_n\setminus x$ converges to the exponential distribution uniformly in
156:   the starting position $x$.
157: \end{theorem}
158: 
159: 
160: In Theorem~\ref{t:sampling} below we state a similar result for
161: another important class of random sets, namely sets $A_n$ that are
162: sampled from $\mathcal V_n$ without replacement (i.e. each subset
163: of $|A_n|$ elements of $\mathcal V_n$ is equally likely).
164: 
165: 
166: \begin{theorem}
167:   \label{t:sampling}
168:   Let $M_n$ be a sequence of integers such that, setting $m(n)={2^n}/{M_n}$,
169:   \begin{equation}
170:     \label{e:condm'}
171:      n \log n \ll m(n) \leq 2^n
172:   \end{equation}
173:   Let $A_n$ be subsets of $M_n$ elements sampled from $\mathcal V_n$ without replacement,
174:   and defined on a common probability space
175:   $(\Omega',F', P')$.
176:   Then, for all $a>0$,
177:   \begin{equation}
178:     \lim_{n\to\infty}
179:     \max_{x\in \mathcal V_n}
180:     \Big |
181:     \mathbb P_x \big[H_n(A_n\setminus x)\ge a m(n)\big]-e^{-a}
182:     \Big |
183:     =0\,,\quad P'\text{-a.s.}
184:   \end{equation}
185: \end{theorem}
186: 
187: 
188: Estimates on the distribution of the hitting time of various subsets of
189: the hypercube have a long history. They can be traced back to the early
190: literature on first passage times for Markov chains (see \cite{Kem61} and
191:   the references therein) where these questions are reformulated in terms
192: of the (one-dimensional) Ehrenfest urn scheme. These results provide very
193: sharp estimates on the asymptotic distribution of the hitting time of a
194: single point. More, recently Matthews \cite[p.118]{Mat89} gave finite-$n$
195: estimates for the Laplace transform of the hitting time of sets containing
196: one or two points.  These estimates are key ingredients that enter his
197: description of the covering time of the hypercube and related questions.
198: 
199: The hitting times of much more general sets (possibly random, and whose
200:   size is possibly increasing with $n$) were studied very recently in
201: \cite{BC06} and \cite{BG06}. In \cite{BG06}, Ben Arous and Gayrard give
202: precise conditions for the hitting time of subsets of $\mathcal V_n$ to be
203: asymptotically exponentially distributed for a class of subsets  for which
204: the so-called lumping construction can be applied.  This construction was
205: fist used in this context by \cite{BBG03}. It can be understood as a $d$-dimensional
206: extension of the Ehrenfest urn scheme where the random walk $Y_n$ on the
207: hypercube is replaced by a walk defined on a $d$-dimensional state space
208: of smaller cardinality, and which evolves in a convex potential that is
209: very steep near its boundary. Such a chain is then studied using the tools
210: developed in \cite{BEGK1},\cite{BEGK2} for the study of metastability in
211: reversible Markov chains on discrete state space. This method allows in
212: particular to show that Theorem~\ref{t:perc} is valid for
213: $m(n)\ge C 2^n/\log n$.
214: 
215: In \cite{BC06}, Ben Arous and \v Cern\'y obtain a result similar to
216: Theorem~\ref{t:perc}. Namely, they prove convergence to the
217: exponential distribution of the hitting time of percolation clouds
218: for densities of order $2^{-c n}$ with $c\in (3/4,1)$ (see
219:   Lemma~3.7 of \cite{BC06}). The method is based on the formula discovered
220: by Matthews in his study of covering problems (see Theorem 1.3 in
221:   \cite{Mat88}) and on (improved) estimates from \cite{Mat89}.
222: 
223:   The results of the present paper are the first that allow to treat sets
224:  of very large size, namely sets as large as $o(2^n (n \log n)^{-1})$.
225: 
226: \smallskip
227: 
228: We will show in Section~\ref{s:equiv} that
229: both Theorem~\ref{t:perc} and Theorem~\ref{t:sampling} are
230: consequences of the following more general theorem. To state it we define
231: the function $\xi_n(k)$ by
232: \begin{equation}
233:   \label{e:defxi}
234:   \xi_n(k)=2^{-n } \frac n 2 \binom n k^{-1}
235:   \sum_{j=1}^{n-k}\binom n {k+j} \frac 1 j,
236:   \qquad k=0,\dots ,n.
237: \end{equation}
238: The role of this relatively complicated function will become evident in
239: Lemma~\ref{l:twopoint}.
240: 
241: \begin{theorem}
242:   \label{t:gen}
243:   Let $n \log n \ll m(n) \leq 2^n$ and
244:   $B_n\subset \mathcal V_n$ be such that
245:   \begin{equation}
246:     \label{e:Bsize}
247:     0<|B_n|= 2^n m(n)^{-1}(1+o(1)).
248:   \end{equation}
249:   Define
250:   \begin{equation}
251:     \begin{split}
252:       \label{e:volumes}
253:       v_n(k)&=\max_{x\in \mathcal V_n} \big|\{y\in B_n:d(x,y)=k\}\big|,\\
254:       V_n(k)&=\max_{x\in \mathcal V_n} \big|\{y\in B_n:d(x,y)\le k\}\big|.
255:     \end{split}
256:   \end{equation}
257:   If there exists a function $g(n)$, such that $g(n)\le n/2$,
258:   \begin{equation}
259:     \label{e:gg}
260:     \xi_n(g(n))\ll 2^{-n}m(n),
261:   \end{equation}
262:   and
263:   \begin{equation}
264:     \label{e:vcond}
265:     \lim_{n\to \infty}\sum_{k=1}^{g(n)-1}v_n(k)\xi_n(k)=0,
266:     \qquad
267:     \text{and}
268:     \qquad
269:     V_n(g(n)-1)\ll |B_n|,
270:   \end{equation}
271:   then
272:   \begin{equation}
273:     \label{e:lim}
274:     \lim_{n\to\infty}
275:     \max_{x\in \mathcal V_n}
276:     \Big |
277:     \mathbb P_x \big[H_n(B_n\setminus x)\ge a m(n)\big]-e^{-a}
278:     \Big |
279:     =0.
280:   \end{equation}
281: \end{theorem}
282: 
283: \begin{remark}
284:   1. While the upper bound on $m(n)$ in the range of validity of the
285:   preceding theorem is trivially optimal, we cannot claim the same about
286:   the lower bound. Morally, the proof exploits the fact that the simple
287:   random walk on the hypercube equilibrates after $O(n \log n)$ steps,
288:   even if an application of this fact is not easy to see in our actual
289:   proof. It is intuitively obvious that after equilibration the hitting
290:   time should be exponentially distributed. The claim of the theorem might
291:   stay valid even for $m(n)=O(n\log n)$, however for different reasons.
292:   Several serious technical complications appears in our proof for such
293:   $m(n)$.  We were not motivated to proceed further since it is not needed
294:   for the applications we have in mind.
295: 
296:   2. The results of  \cite{BG06} (see Theorem 1.7 of \cite{BG06} and the
297:     simpler Corollary 1.8) imply that Theorem~\ref{t:perc} holds true for
298:   sets $B_n$ whose size is asymptotically constant that is, when
299:   \eqref{e:Bsize} holds for $m(n)={2^n}/{M}$ where $M$ is a fixed integer.
300: 
301:   3. We will show soon that $\xi_n(1)=O(n^{-1})$. Therefore, the
302:   conditions appearing in \eqref{e:vcond} are void when \eqref{e:Bsize}
303:   holds with $m(n)$ satisfying $2^{-n}m(n)\gg n^{-1}$ (and, in
304:     particular, when $|B_n|$ is constant).  As an immediate consequence of
305:   Theorem~\ref{t:gen} we then obtain:
306: \end{remark}
307: 
308: \begin{corollary}
309:   \label{c:perc}
310:   Theorem~\ref{t:perc} remains valid on $2^{n}(\log n)^{-1}\le \bar m(n)\leq
311:   2^n$ if the term $a\bar m(n)$ in \eqref{e:thmjedna} is replaced by $2^n |A_n|^{-1}$.
312: \end{corollary}
313: 
314: \begin{proof}
315:   It is sufficient to take $m(n)$ in Theorem~\ref{t:gen}
316:   to be $m(n)=2^n/|A_n|$, where $A_n$ is the percolation cloud with
317:   intensity $\bar m(n)$.
318: \end{proof}
319: 
320: Our method of proof relies on a sharp estimate on the distribution
321: of hitting time of a single point when time is measured on the
322: scale $m(n)$, namely on the quantity
323: \begin{equation}
324:   \mathbb P_{z_k}[H_n(\boldsymbol 0)< a m(n)]
325: \end{equation}
326: where $\boldsymbol 0$ is the vertex of $\mathcal V_n$ whose coordinates
327: are all $0$ and $z_k$ is any vertex at distance $k$ from it; this estimate
328: is itself derived from an estimate on the Laplace transform of the hitting
329: time $H_n(\boldsymbol 0)$ (see respectively Lemma~\ref{l:twopoint} and
330:   Lemma~\ref{l:laplace}). The domain of validity of the latter determines
331: our bounds on $m(n)$.  On this domain the Laplace transform is well
332: approximated by the sum of two terms:  the expected contribution of an
333: exponential distribution, and the mysterious looking function $\xi_n(k)$
334: which is, essentially, the probability that $Y_n$ started from $z_k$ hits
335: $\boldsymbol 0$ in the first $n\log n$ steps.
336: 
337: The rest of the paper is organised as follows. In the next section we
338: describe our main motivation and give some important consequences of our
339: results for aging in the Random Energy model.
340: In Section~\ref{s:prop} we give the proof of Theorem~\ref{t:gen}.
341: Finally, Theorems~\ref{t:perc} and~\ref{t:sampling} are proved in
342: Section~\ref{s:equiv}.
343: %%%% END OF ht1.tex %%%%
344: 
345: %%%% START OF ht4.tex %%%%
346: 
347: \section{Aging in the Random Energy Model}
348: \label{s:aging}
349: 
350: The main motivation behind this paper originates in the study of the
351: Random Hopping Time (RHT) dynamics of the Random Energy Model (REM), which
352: is often called the simplest model of a spin glass. Let us describe this problem
353: briefly (for a recent review  see \cite{BC06c}). In the REM an energy $E_x$
354: is attached to every site $x\in \mathcal V_n$. The $E_x$'s are chosen to
355: be i.i.d.~with standard normal distribution. Given a collection
356: $\boldsymbol E=\{E_x:x\in \mathcal V_n\}$ and a parameter $\beta>0 $
357: (representing the inverse of the temperature), the RHT dynamics in the REM
358: is defined as a continuous-time Markov process $X_n(\,\cdot\,)$ whose
359: transition rates are given by
360: $w^{\boldsymbol E}_n(x,y)=\exp(-\beta \sqrt n E_x)$ if $d(x,y)=1$ and zero
361: otherwise.
362: 
363: The main goal of the study of the processes $X_n$ was to prove aging. In
364: this context it usually means  showing that
365: the  two-point function
366: \begin{equation}
367:   R_n(t_w,t_w+t;\boldsymbol E):=P[X_n(t_w+t)=X_n(t_w)|\boldsymbol E],
368: \end{equation}
369: has a non-trivial limit as $t_w$, $ t=\theta t_w$ and $n$ tend simultaneously
370: to infinity.
371: 
372: The first proof of aging in the REM for $\beta > \beta_c=\sqrt {2 \log 2}$
373: and times $t_w\sim \exp (\beta \beta_c n)$ was given in
374: \cite{BBG03,BBG03b}, based on the analysis of the metastable behaviour of
375: $X_n$ and renewal theory. In \cite{BC06} another proof
376: of aging, based on the arc-sine law for stable subordinators, was given
377: for temperatures satisfying
378: \begin{equation}
379:   \label{e:bc}
380:   \sqrt {3/4} < \alpha \beta / \beta_c < 1.
381: \end{equation}
382: and (shorter) time-scales
383: $t_w(n)=( \alpha \beta \sqrt{2 \pi n})^{-1/\alpha }\exp(\alpha \beta^2 n)$,
384: where $\alpha \in (0,1)$ is a free parameter. We will now explain how our
385: methods allows to improve the lower bound $\sqrt{3/4}$ in \eqref{e:bc} to its
386: optimal value, that is to $0$. To this end we need to describe briefly the
387: background of the techniques of \cite{BC06}.
388: 
389: The transition rates of the process $X$ do not depend on the energy of the
390: target vertex $y$. Therefore, the process $X_n$ is  a time change of the simple
391: random walk $Y_n$, and  can be written as
392: $X_n(t)=Y_n(S^{-1}_n(t))$, where
393: \begin{equation}
394:   S_n(k)=\sum_{i=0}^{k-1} e_i \exp(\beta \sqrt n E_{Y_n(i)}),
395: \end{equation}
396: $S^{-1}_n$ is the generalised right-continuous inverse of $S_n(k)$, and
397: $e_i$ is a sequence of mean-one i.i.d.~exponential random variables.
398: $S_n(k)$ is the time of the $k^{\text {th}}$ jump of $X_n$.
399: 
400: It was proved in \cite{BC06} that $S(n)$ behaves (for $n$ large) as an
401: $\alpha $-stable subordinator  in certain time and temperature regimes. To
402: be more precise, set $r(n)=\exp(\alpha^2 \beta^2 n /2)$. Then, if
403: \eqref{e:bc} holds then   $t(n)^{-1}S_n(\lfloor  r(n) \,\cdot\,\rfloor)$
404: converges in distribution to an $\alpha $-stable subordinator.
405: 
406: It is a known fact that the value of a stable subordinator at time $t$ can
407: be approximated by the finite sum of its largest jumps up to this time.
408: The same is true for $S_n$: the main contribution to $S_n(r(n))$ comes
409: from a finite number of visits to sites with
410: $\exp(\beta \sqrt n E_x)\asymp t(n)$. Such sites  form (due to the
411:   i.i.d.~property of the energies) a percolation cloud.  To understand
412: properties of $S_n$ it is therefore necessary to understand how the simple
413: random walk visits such clouds.
414: 
415: As we have already remarked, the methods of \cite{BC06} are sufficient to
416: show Theorem~\ref{t:perc} for densities much smaller than $2^{-3n/4}$.
417: The constant $3/4$ in the exponent entails the constant $3/4$ in \eqref{e:bc}. The
418: result of our Theorem~\ref{t:perc} thus allows to extend the domain of
419: validity of Theorem 3.1 in \cite{BC06}.  As we find this extension important
420: we state it here:
421: 
422: \begin{theorem}
423:   For $\alpha \in (0,1)$ let $t_w(n)$ and $r(n)$ be as above. If
424:   \begin{equation}
425:     \label{e:new}
426:     0 < \alpha \beta < \beta_c,
427:   \end{equation}
428:   then $t_w(n)^{-1}S_n(r(n)\,\cdot\,)$ converges in distribution to an
429:   $\alpha $-stable subordinator and the two-point function $R_n$ exhibits
430:   aging. Namely, for a.e.~realisation of
431:   $\boldsymbol E$ and for every $\theta\in (0,\infty)$
432:   \def\Asl{\mathop{\mathsf{Asl}}\nolimits}
433:   \begin{equation}
434:     \lim_{n\to\infty} R_n(t_w(n),(1+\theta ) t_w(n);\boldsymbol E)=\Asl_\alpha (1/1+\theta ),
435:   \end{equation}
436:   where $\Asl_\alpha (u)$ stands for the distribution function of the
437:   generalised arcsine law with parameter $\alpha $,
438:   $ \Asl_\alpha (z):= \pi^{-1}\sin \alpha \pi  \int_0^{z}u^{\alpha
439:       -1}(1-u)^{-\alpha }\,\d u$.
440: \end{theorem}
441: 
442: It is worth noting that for any $\beta<\infty $ there exists $\alpha \in (0,1)$
443: such that \eqref{e:new} is satisfied. This implies that aging occurs in the
444: RHT dynamics of the REM at all temperatures.
445: 
446: \smallskip
447: 
448: Theorem~\ref{t:perc} allows further to study the RHT dynamics in
449: another interesting time-temperature regime where denser percolation
450: clouds should be considered:  for
451: $t_w(n)=(\beta \sqrt{2 \pi n})^{-1}\exp(\beta^2 n)$, that is for
452: $\alpha =1$. It is argued in the physics literature \cite{BB02b}
453: that the two-point function $R_n$ exhibits some interesting ultrametric
454: behaviour in this case. The rigorous treatment of this problem was the
455: original motivation behind this paper and will be subject of
456: a forthcoming paper.
457: 
458: 
459: 
460: %%%% END OF ht4.tex %%%%
461: 
462: %%%% START OF ht2.tex %%%%
463: 
464: \section{Proof of Theorem~\ref{t:gen}}
465: \label{s:prop}
466: 
467: As mentioned earlier Theorem~\ref{t:gen} is already known for small sets,  namely when
468: $m(n)={2^n}/{M}$ for $M$ a fixed integer (see \cite{BG06}, Corollary 1.8). Although
469: our method of proof clearly allows us to cover this case as well,
470: its treatment is, in places, quite different from the case $n \log
471: n \ll m(n) \ll 2^n$. Thus, in order to keep this paper as concise as
472: possible we will prove Theorem~\ref{t:gen} in that
473: latter case only and assume from now on that $n \log n \ll
474: m(n) \ll 2^n$.
475: 
476: \begin{proof}[Proof of Theorem~\ref{t:gen} (for $n \log n \ll m(n) \ll 2^n$)]
477:   We have
478:   \begin{equation}
479:     \mathbb P_x[H_n(B_n\setminus x)\ge a m(n)]
480:     =
481:     1- \mathbb P_x\Big[\bigcup_{y\in B_n\setminus x}
482:       H_n(y)< a m(n)\Big].
483:   \end{equation}
484:   Therefore,
485:   it follows from the inclusion-exclusion principle that
486:   for all even $\ell\in \mathbb N$
487:   \begin{equation}
488:     \mathbb P_x[H_n(B_n\setminus x)\ge a m(n)]
489:     \le
490:     1+\sum_{i=1}^\ell \frac {(-1)^i} {i!}
491:     \psum_{y_1,\dots,y_i\in B_n\setminus x}
492:     \mathbb P_x\Big[\bigcap_{j=1}^i H_n(y_j)< a m(n)\Big],
493:   \end{equation}
494:   where $\psum$ denotes the sum over all mutually different $y$'s.
495:   Analogous expressions for~$\ell$ odd give lower bounds.
496: 
497:   The following proposition is the
498:   key step of the proof.
499:   \begin{proposition}
500:     \label{p:sum}
501:     Under the assumptions of Theorem~\ref{t:gen}
502:     and assuming that $n \log n \ll m(n) \ll 2^n$ we have,
503:     for all $i\in \mathbb N$,
504:     \begin{equation}
505:       \lim_{n\to\infty}
506:       \max_{x\in \mathcal V_n}
507:       \bigg|
508:       \psum_{y_1,\dots,y_i\in B_n\setminus x}
509:       \mathbb P_x\Big[\bigcap_{j=1}^i H_n(y_j)< a m(n)\Big]
510:       -a^i\bigg| =0.
511:     \end{equation}
512:   \end{proposition}
513:   Using Proposition~\ref{p:sum} the completion of the proof of Theorem~\ref{t:gen}
514:   under the assumption that $n \log n \ll m(n) \ll 2^n$ is immediate.
515: \end{proof}
516: 
517: 
518: The proof of Proposition~\ref{p:sum} relies on several technical
519: lemmas which we collect in the subsection below. The proof of
520: Proposition~\ref{p:sum} is then concluded in subsection
521: \ref{s:propproved}.
522: 
523: \subsection{Preparatory Lemmas}
524: \label{s:preplem}
525: 
526: Our first lemma collects the properties of the function $\xi_n(k)$
527: which will be needed later.
528: 
529: \begin{lemma}
530:   \label{l:xi}
531:   (i) For all $k\in \{1,\dots,n\}$,
532:   $ \xi_n(k) \le K  \binom n k^{-1} n^{1/2}\log n$, where $K$ is a constant
533:   independent of $n$ and $k$.
534: 
535:   (ii) For all  $k\le n/2$,  $ \xi_n(k)\ge    \frac 1 2 \binom n k^{-1}$.
536: 
537:   (iii) For any fixed $n$ the function $\xi_n(k)$ is decreasing in $k$.
538: 
539:   (iv) If $k=o(n)$, then $ \xi_n(k)=\binom n k^{-1}(1+o(1))$.
540: \end{lemma}
541: \begin{proof}
542:   (i) Recall that
543:   $\xi_n(k)= 2^{-n}\frac n 2 \binom n k^{-1}\sum_{j=1}^{n-k}\binom n {j+k} \frac 1 j$.
544:   From a standard moderate deviations argument it follows that
545:   \begin{equation}
546:     2^{-n} \frac n 2 \sum_{\substack{j=0\\|j-n/2|\ge n^{7/12}}}^n \binom n j
547:     \le c_1 \frac n 2 e^{-c_2 n^{1/6}}.
548:   \end{equation}
549:   Therefore, the contribution of $j$'s with $|k+j-n/2|\ge n^{7/12}$ is
550:   $o(\binom  n k^{-1})$. For the remaining $j$'s we use the approximation
551:   \begin{equation}
552:     \label{e:binomass}
553:     \binom n {n/2+i} = \sqrt{\frac 2 \pi } n^{-1/2}2^n
554:     e^{-2 i^2/n}(1+o(1)),
555:   \end{equation}
556:   which is valid uniformly for $i=o(n^{2/3})$. Setting
557:   $a=n^{-1/2}(k+j-n/2)$ and $b=n^{-1/2}(k-n/2)$, and thus $j=n^{1/2}(a-b)$,
558:   the  contribution of the $j$'s with $|k+j-n/2|\le n^{7/12}$ to
559:   $\xi_n(k)$ equals
560:   \begin{equation}
561:     \begin{split}
562:       \label{e:gausapprox}
563:       2^{-n}\frac n2&
564:       \binom  n k^{-1}
565:       \sum_{\substack{a\in [-n^{1/12},n^{1/12}]\cap (\mathbb Z/\sqrt n)\\
566:           a\ge b+n^{-1/2}}}
567:       \binom n {a \sqrt n + n/2} \frac 1 {(a-b)\sqrt n}
568:       \\&\le K'
569:        n^{1/2}
570:       \binom  n k^{-1}
571:        \int_{( b+ n^{-1/2})\vee-n^{1/12}}^{n^{1/12}}
572:       e^{-2x^2}\frac {\d x} {(x-b)}
573:       \le
574:       K
575:       \binom  n k^{-1}
576:       n^{1/2}\log n.
577:     \end{split}
578:   \end{equation}
579: 
580:   (ii) For $k<n/2$,
581:   $\xi_n(k)\ge 2^{-n}\frac n 2 \binom n k^{-1} \sum_{j=1}^{n/2}\binom n {k+j}\frac 2 n
582:   \ge \frac 1 2 \binom n k^{-1}$.
583: 
584:   (iii) The function $\xi_n$ can be rewritten as
585:   \begin{equation}
586:     \label{e:secondform}
587:     \xi_n(k)=
588:     2^{-n}\frac n2
589:     \sum_{j=1}^{n-k}
590:     \binom{n-k} j
591:     \binom{k+j} j^{-1}
592:     \frac 1 j .
593:   \end{equation}
594:   Here, the fact that $\xi_n(k)$ is decreasing is apparent.
595: 
596: 
597:   (iv)  Using again the moderate  deviations argument, for $k=o(n)$,
598:   \begin{equation}
599:     \xi_n(k)=2^{-n}\frac n 2\binom n k^{-1}
600:     \sum_{j=n/2-n^{7/12}-k}^{n/2+n^{7/12}-k}\binom n {j+k} \frac
601:     1 j(1+o(1))=\binom n k^{-1}(1+o(1)).
602:   \end{equation}
603:   This completes the proof of Lemma~\ref{l:xi}.
604: \end{proof}
605: 
606: We now prove that for $m(n)$ in the range considered in Theorem~\ref{t:gen}
607: there always exists a function $g(n)$ with $g(n)\le n/2$ satisfying \eqref{e:gg}.
608: We in fact prove a little more:
609: \begin{lemma}
610:   Let $m(n)$ be such that $n \log n \ll m(n)\ll 2^n$. Then there exist
611:   function $g$ such that
612:   \begin{equation}
613:     \label{e:g}
614:     n/2-g(n)\gg n^{1/2}
615:     \qquad\text{and}\qquad
616:     \xi_n(g(n))\ll 2^{-n}m(n).
617:   \end{equation}
618: \end{lemma}
619: \begin{proof}
620:   Take $m'(n)$ such that $n\log n \ll m'(n)\ll m(n)$ and define $g(n)$ by
621:   \begin{equation}
622:     g(n)=\min\{ k: \xi_n(k)\le 2^{-n}m'(n)\}.
623:   \end{equation}
624:   Such $g(n)$ satisfies the second half of \eqref{e:g} by definition.
625:   By Lemma~\ref{l:xi}(i)
626:   \begin{equation}
627:     g(n)\le
628:     \min\Big\{k: K \binom n k^{-1} n^{1/2} \log n\le 2^{-n}m'(n)\Big\}
629:   \end{equation}
630:   Write $k=n/2-i$. Using formula \eqref{e:binomass} for the binomial
631:   coefficient we find that
632:   \begin{equation}
633:     g(n)\le \frac n2
634:     +\min\Big\{i: e^{2i^2/n}\le \frac {c m'(n)}{n\log n}\Big\}.
635:   \end{equation}
636:   Since $m'(n)\gg n\log n$ there exists $i_0$, $n^{2/3}\gg i_0\gg n^{1/2}$, for
637:   which the inequality in braces holds. Since $n/2-g(n)\ge i_0$, the first
638:   half of \eqref{e:g} is proved.
639: \end{proof}
640: 
641: 
642: Let us assume from now on that $g(n)$ satisfies \eqref{e:g}. As
643: announced earlier, the key ingredient of the proof of
644: Proposition~\ref{p:sum} is a precise estimate on the Laplace
645: transform of the hitting time of a single point. We now state and
646: prove this result. As we will see, this is where the function
647: $\xi_n$ comes in.
648: 
649: 
650: \begin{lemma}
651:   \label{l:laplace}
652:   Let $\boldsymbol 0$ be the vertex
653:   of the hypercube with all coordinates equal to $0$ and
654:   let $z_k\in \mathcal V_n$ be an arbitrary
655:   vertex of the hypercube such that $d(z_k,\boldsymbol 0)=k$. If
656:   $n\log n \ll m(n) \ll 2^n$, then for all $s>0$
657:   \begin{equation}
658:     \mathbb E_{z_k}\exp\Big(-\frac s {m(n)} H_n(\boldsymbol 0)\Big)=
659:     \Big[2^{-n}\frac {m(n)} s + \xi_n(k)\Big](1+o(1)).
660:   \end{equation}
661: \end{lemma}
662: \begin{proof}
663:   By Fourier methods for random walks on finite groups \cite{Dia88}, we
664:   have as in \cite{Mat89,BC06}
665:   \begin{equation}
666:     \label{e:furt}
667:     \mathbb E_{z_k} e^{-s/m H_n(\boldsymbol 0)}=
668:     \frac
669:     {\sum_{y\in \mathcal V_n}(-1)^{z_k\cdot y}\big[1-e^{-s/m }
670:     (1-\frac {2 d(y,\boldsymbol 0)}{n})\big]^{-1}}
671:     {\sum_{y\in \mathcal V_n}\big[1-e^{-s/m}
672:     (1-\frac {2 d(y,\boldsymbol 0)}{n})\big]^{-1}},
673:   \end{equation}
674:   where $x\cdot y = \sum_{i=1}^n x(i) y(i)$ is the standard scalar
675:   product in $\mathbb R^n$.
676: 
677:   Let us first consider the numerator of \eqref{e:furt}. Observe that there are
678:   $\tbinom k i \tbinom {n-k} j$ sites $y\in \mathcal V_n$ such that
679:   $d(0,y)=i+j$ and $z_k\cdot y = i$. Hence the numerator of \eqref{e:furt} is equal to
680:   \begin{equation}
681:     \sum_{i=0}^k \sum_{j=0}^{n-k}
682:     (-1)^i \binom k i \binom {n-k} j
683:     \frac 1 {1-e^{-s/m}(1-2n^{-1}(i+j))}.
684:   \end{equation}
685:   This expression can be simplified using the following lemma.
686:   \begin{lemma}
687:     For all $k,j\in \{0,1,\dots\}$ and all $s>0$
688:     \begin{equation}
689:       \label{e:betaint}
690:       \sum_{i=0}^k
691:       \frac{(-1)^i \binom k i}
692:       {1-e^{-s/m}(1-2n^{-1}(i+j))}
693:       =
694:       \frac {ne^{s/m}} 2 \cdot
695:       \frac{\Gamma (1+k)\Gamma(j+\frac n2(e^{s/m}-1))}
696:       {\Gamma(1+k+j+\frac n2 (e^{s/m}-1))}
697:       .
698:     \end{equation}
699:   \end{lemma}
700:   \begin{proof}
701:     Note that the second fraction on the right-hand side of
702:     \eqref{e:betaint} can be expressed using the Beta-integral,
703:     \begin{equation}
704:       \frac {ne^{s/m}} 2 \cdot
705:       \frac{\Gamma (1+k)\Gamma(j+\frac n2(e^{s/m}-1))}
706:       {\Gamma(1+k+j+\frac n2 (e^{s/m}-1))}
707:       =
708:       \frac {ne^{s/m}} 2
709:       \int_0^1
710:       (1-t)^k t^{i+\frac n2 (e^{s/m}-1)-1}\,\d t.
711:     \end{equation}
712:     Expanding $(1-t)^k$ according to the binomial theorem, and performing an easy
713:     integration then gives the left-hand side of \eqref{e:betaint}.
714:   \end{proof}
715: 
716:   Using the last lemma the numerator of \eqref{e:furt} can be rewritten as
717:   \begin{equation}
718:     \label{e:num2}
719:     \sum_{j=0}^{n-k}
720:     \binom {n-k} j
721:     \frac {ne^{s/m}} 2 \cdot
722:     \frac{\Gamma (1+k)\Gamma(j+\frac n2(e^{s/m}-1))}
723:     {\Gamma(1+k+j+\frac n2 (e^{s/m}-1))}.
724:   \end{equation}
725: 
726:   So far we obtained an exact expression which we now want to evaluate. To do so
727:   we will use the following known properties of $\Gamma $-functions
728:   (we refer to \cite{AS72} for the definition and the properties of the functions
729:   appearing below).
730:   \begin{equation}
731:     \Gamma '(x)=\Gamma (x) \psi _0(x),\qquad
732:     \Gamma''(x)=\Gamma (x)(\psi _0(x)^2+\psi _1(x)),\qquad
733:     \lim_{x\to 0}x\Gamma (x)=1,
734:   \end{equation}
735:   where $\psi _0$ is the digamma function and $\psi _1=\psi_0'$.
736:   The values of these functions for integer arguments can be written explicitly:
737:   \begin{equation}
738:     \psi _0(k)=\sum_{i=1}^{k-1} \frac 1 i - \gamma_E = \log k(1+o(1)), \qquad
739:     \psi _1(k)=\frac {\pi^2} 6- \sum_{i=1}^{k-1} \frac 1 {i^2}=O(1/k).
740:   \end{equation}
741:   where $\gamma_E$ is Euler's constant.
742: 
743:   We can now evaluate \eqref{e:num2}. Set $\varepsilon = \frac n2(e^{s/m}-1)$
744:   and observe that the bound $m\gg n\log n$ entails $\varepsilon \ll (\log n)^{-1}$.
745:   To treat the term $j=0$ in \eqref{e:num2} simply note that,
746:   since $\varepsilon \psi_0(1+k)=o(1)$ for all $k\le n$,
747:   \begin{equation}
748:     \label{e:blams}
749:     \frac {ne^{s/m}} 2 \cdot
750:     \frac{\Gamma (1+k)\Gamma(\varepsilon )}
751:     {\Gamma(1+k+\varepsilon )}
752:     =
753:     \frac {ne^{s/m}} 2 \cdot
754:     \frac {\varepsilon^{-1}(1+o(1))}
755:     {1+\varepsilon \psi _0(1+k)}=\frac m s
756:     (1+o(1)).
757:   \end{equation}
758:   A similar calculation for the remaining terms combined with
759:   \eqref{e:secondform} readily yields
760:   \begin{equation}
761:     \begin{split}
762:       \label{e:blaxi}
763:       \sum_{j=1}^{n-k}&
764:       \binom {n-k} j
765:       \frac {ne^{s/m}} 2 \cdot
766:       \frac{\Gamma (1+k)\Gamma(j+\varepsilon )}
767:       {\Gamma(1+k+j+\varepsilon )}
768:       \\&=
769:       \sum_{j=1}^{n-k}
770:       \frac n {2j} \binom {n-k} j {\binom{k+j} j}^{-1}
771:       \frac {(1+\varepsilon \psi _0(j))(1+o(1))}
772:       {1+\varepsilon \psi _0(1+k+j)}
773:       =2^n \xi_n(k)(1+o(1)).
774:     \end{split}
775:   \end{equation}
776: 
777:   For $k=0$ the numerator of \eqref{e:furt} coincides with the denominator.
778:   Equations \eqref{e:blams}, \eqref{e:blaxi} and Lemma~\ref{l:xi}(iv)
779:   then imply that the denominator behaves like
780:   \begin{equation}
781:     \label{e:bladem}
782:     \{ms^{-1} + 2^n \xi_n(0)\}(1+o(1))=2^n(1+o(1)).
783:   \end{equation}
784:   Finally, putting together \eqref{e:blams}, \eqref{e:blaxi} and \eqref{e:bladem}
785:   yields the claim of Lemma~\ref{l:laplace}.
786: \end{proof}
787: 
788: 
789: Lemma~\ref{l:laplace} now allows us to get information on the form
790: of the probability distribution function of $H_n(\boldsymbol 0)$.
791: Let us denote by $p_n(a,k)$ the probability
792: \begin{equation}
793:   p_n(a,k)=\mathbb P_{z_k}[H_n(\boldsymbol 0)< a m(n)].
794: \end{equation}
795: 
796: \begin{lemma}
797:   \label{l:twopoint}
798:   (i) There exists $C<\infty$ independent of $n$ and $k$ such that
799:   \begin{equation}
800:     \label{e:pnupperbound}
801:     p_n(a,k)\le Ce^a\big(2^{-n}m(n) +  \xi_n(k)\big).
802:   \end{equation}
803: 
804:   (ii)  For
805:   any $a\in [0,\infty)$, uniformly on compact subsets of this interval,
806:   \begin{equation}
807:     \lim_{n\to\infty}
808:     \max_{k\ge g(n)}
809:     \big |2^n m(n)^{-1}p_n(a,k)-a\big|=0.
810:   \end{equation}
811: \end{lemma}
812: 
813: \begin{proof}
814:   Assertion (i)  follows from Chebyshev
815:   inequality and Lemma~\ref{l:laplace}. To prove (ii) observe
816:   that
817:   for $k\ge g(n)$, by \eqref{e:g} and by Lemma~\ref{l:xi}(iii),
818:   $\xi_n(k)\le \xi_n(g(n))\ll 2^{-n}m(n)$. Therefore,
819:   $\mathbb E_{z_k}\exp(-s H_n(\boldsymbol 0)/m(n))=2^{-n}m(n)/s(1+o(1))$.
820:   Consider the sequence of measures $\mu_n$ given by
821:   \begin{equation}
822:     \mu_n([0,t])= 2^n m(n)^{-1} \mathbb P_{z_k}\big[H_n(\boldsymbol 0)/m(n)\in
823:       [0,t]\big].
824:   \end{equation}
825:   The Laplace transform of $\mu_n$ then satisfies
826:   $  \int_0^\infty e^{-s t}\mu_n(\d t)\xrightarrow{n\to\infty}  1/s$.
827:   Therefore, $\mu_n$ converges weakly to the Lebesgue measure (see
828:     \cite{Fel71},  Section XIII.1, Theorem 2a, p. 433) and thus
829:   \begin{equation}
830:     2^n m(n)^{-1}p_n(a,k)=\mu_n([0,a))\xrightarrow{n\to\infty} a.
831:   \end{equation}
832:   The uniformity on compact sets follows easily from the fact that
833:   the probabilities $p_n(a,k)$ are increasing in $a$.
834: \end{proof}
835: 
836: We finally use Lemma~\ref{l:twopoint} to get information on the
837: form of the probability distribution function of the hitting time
838: of finite subsets of points of $\mathcal V_n$. For
839: $y_1,\dots,y_\ell\in \mathcal V_n$ we define $\bar
840: H_n(y_1,\dots,y_\ell)$ by
841: \begin{equation}
842:   \bar H_n(y_1,\dots,y_\ell)=
843:   \begin{cases}
844:     H_n(y_\ell)&\text{if $H_n(y_i)<H_n(y_{i+1})$ for all
845:       $i\in \{1,\dots,\ell -1\}$},\\
846:     \infty&\text{otherwise}.
847:   \end{cases}
848: \end{equation}
849: That is $\bar H_n(y_1,\dots,y_\ell)$ is finite only if the $y$'s were
850: visited in the prescribed order. In this case it is equal to the time to
851: visit all $y$'s. Observe that it is always infinite if $y_i=y_j$ for some
852: $i\neq j$.
853: 
854: \begin{lemma}
855:   \label{l:coverbad}
856:   Let $x=y_0$, let $y_1,\dots,y_\ell$ be mutually distinct points in
857:   $\mathcal V_n$, and let $d(i)=d(y_{i-1},y_i)$. Then
858:   \begin{equation}
859:     \label{e:coverbad}
860:     \mathbb P_x\big[\bar H_n(y_1,\dots,y_\ell)<am(n)\big]
861:     \le
862:     C^\ell e^{\ell a} \prod_{i=1}^\ell \big(2^{-n}m(n)+ \xi_n(d(i))\big).
863:   \end{equation}
864: \end{lemma}
865: 
866: \begin{proof}
867:   Obviously, by the strong Markov property,
868:   \begin{equation}
869:     \mathbb P_x\big[\bar H_n(y_1,\dots,y_\ell)<am(n)\big] \le
870:     \prod_{i=1}^\ell
871:     \mathbb P_{y_{i-1}}[H_n(y_i)< am(n)],
872:   \end{equation}
873:   which, by Lemma~\ref{l:twopoint}(i), is bounded by the right-hand side of
874:   \eqref{e:coverbad}.
875: \end{proof}
876: 
877: \begin{lemma}
878:   \label{l:covergood}
879:   Let $x^n=y_0^n$, let $y_1^n,\dots,y_\ell^n\in \mathcal V_n$, and let
880:   $d_n(i)=d(y^n_i,y^n_{i-1})$. Suppose that $d_n(i)\ge g(n)$ for
881:   all $i\in\{1,\dots,\ell\}$ and all $n$.
882:   Then, uniformly over all $x^n$ and $y^n_i$,
883:   \begin{equation}
884:     \lim_{n\to\infty}
885:     2^{\ell n} m(n)^{-\ell}
886:     \mathbb P_{x^n}\big[\bar H_n(y_1^n,\dots,y_\ell^n)
887:       < a m(n)\big] =
888:     \frac {a^\ell}{\ell !}.
889:   \end{equation}
890: \end{lemma}
891: 
892: \begin{proof}
893:   The probability in question can be bounded from above using the strong Markov
894:   property,
895:   \begin{equation}
896:     \label{e:iiint}
897:     \mathbb P_{x^n}\big[\bar H_n(y_1^n,\dots,y_\ell^n)
898:       < a m(n)\big] \le
899:     \sum_{\substack{a_1,\dots, a_\ell\in \mathbb N/m(n)
900:         \\ a_1+\dots+a_\ell < a}}
901:     \prod_{i=1}^\ell
902:     \mathbb P_{y_{i-1}}[H_n(y_i)/m(n)= a_i].
903:   \end{equation}
904:   Since $d(y_{i-1},y_i)>g(n)$, it is easy to
905:   see from Lemma~\ref{l:twopoint}(ii) that the sum behaves like
906:   \begin{equation}
907:     2^{-n\ell} m(n)^{\ell}
908:     \idotsint\limits_{x_1+\dots+x_\ell < a}
909:     \d x_1\dots\d x_\ell
910:     (1+o(1))
911:     =
912:     2^{-n\ell} m(n)^{\ell}
913:     a^\ell / \ell !
914:     (1+o(1)).
915:   \end{equation}
916: 
917:   The above expression only provides an upper bound since it does not
918:   exclude the possibility that the random walk visits another $y_j$, $j>i$, on its way from
919:   $y_{i-1}$ to $y_i$.
920:   To construct a lower bound we should exclude such
921:   visits. Therefore, denoting by $\mathrm{UB}$ the upper
922:   bound~\eqref{e:iiint},
923:   \begin{equation}
924:     \begin{split}
925:       \mathbb P_{x^n}&\big[\bar H_n(y_1^n,\dots,y_\ell^n)
926:         < a m(n)\big]
927:       \\
928:       &\ge \mathrm{UB}-
929:       \sum_{i=1}^\ell
930:       \sum_{j=i+1}^\ell
931:       \sum_{\substack{a_1,\dots, a_\ell\in \mathbb N/m(n)
932:           \\ a_1+\dots+a_\ell < a}}
933:       \prod_{\substack{k=1\\k\neq i}}^\ell
934:       \mathbb P_{y_{k-1}}[H_n(y_k)/m(n)= a_k]
935:       \\&\quad\times
936:       \sum_{\substack{b\in \mathbb N/m(n)\\ b\in(0,a_i-a_{i-1})}}
937:       \mathbb P_{y_{i-1}}[H_n(y_j)/m(n)= b]
938:       \mathbb P_{y_{j}}[H_n(y_i)/m(n)= a_i-b].
939:     \end{split}
940:   \end{equation}
941:   The negative term on the right-hand side is smaller than
942:   \begin{equation}
943:     C2^{-n(\ell-1)} m(n)^{\ell-1}
944:     e^{(\ell-1)a}
945:     \sum_{i=1}^\ell
946:     \sum_{j=i+1}^\ell
947:     \mathbb P_{y_{i-1}}[H_n(y_j)< am(n)]
948:     \mathbb P_{y_{j}}[H_n(y_i)<am(n) ].
949:   \end{equation}
950:   Hence, if we can show that for all $j\neq i$
951:   \begin{equation}
952:     \label{e:smalll}
953:     \mathbb P_{y_{i-1}}[H_n(y_j)< am(n)]
954:     \mathbb P_{y_{j}}[H_n(y_i)<am(n) ] = o(2^{-n}m(n)),
955:   \end{equation}
956:   then the proof of Lemma~\ref{l:covergood} is finished.
957: 
958:   Let $k=d(y_{i-1},y_j)$ and $l=d(y_j,y_i)$.
959:   Since $d(y_{i-1},y_i)\ge g(n)$ we have also $k+l\ge g(n)$. Obviously
960:   $k\ge 1$, $l\ge 1$. By Lemma~\ref{l:twopoint}(i),
961:   \begin{equation}
962:     \begin{split}
963:       \mathbb P_{y_{i-1}}&[H_n(y_j)< am(n)]
964:       \mathbb P_{y_{j}}[H_n(y_i)<am(n) ]
965:       \\&\le
966:       C^2e^{2a}\big(2^{-n}m(n)+ \xi_n(k)\big).
967:       \big(2^{-n}m(n)+ \xi_n(l)\big)
968:       \\&\le
969:       C^2e^{2a}\big(2^{-2n}m(n)^{2}+ 2^{-n}m(n) 2\xi_n(1)+\xi_n(k)\xi_n(l)\big).
970:     \end{split}
971:   \end{equation}
972:   The first two summands are $o(2^{-n}m(n))$ (see Lemma~\ref{l:xi}(iii,iv)).
973:   If $k$ or $l$ is larger than $g(n)$, then the same is valid for the
974:   third one. As the last step of the proof we show that if
975:   $\max \{k,l\}< g (n)$ and $k+l\ge g(n)$, then for any $\varepsilon >0$ and
976:   $n$ large enough
977:   \begin{equation}
978:     \label{e:submult}
979:     \xi_n(k)\xi_n(l)\le \varepsilon 2^{-n} m(n).
980:   \end{equation}
981:   Let $z_{k+l}$ be as in Lemma~\ref{l:laplace}
982:   and let $z_k$ be any point such that $d(z_k,\boldsymbol 0)=k$ and
983:   $d(z_k,z_{k+l})=l$. Since on the way from $z_{k+l}$ to $\boldsymbol 0$,
984:   the random walk may pass through $z_k$ we have
985:   \begin{equation}
986:     \mathbb E_{z_{k+l}}[e^{-sH_n(\boldsymbol 0)}]
987:     \ge
988:     \mathbb E_{z_{k+l}}[e^{-sH_n(z_k)}]
989:     \mathbb E_{z_{k}}[e^{-sH_n(\boldsymbol 0)}].
990:   \end{equation}
991:   Lemma~\ref{l:laplace} then yields
992:   \begin{equation}
993:     \big\{2^{-n}\frac {m(n)} s + \xi_n(k+l)\big\}(1+o(1))
994:     \ge
995:     \big\{2^{-n}\frac {m(n)} s + \xi_n(k)\big\}
996:     \big\{2^{-n}\frac {m(n)} s + \xi_n(l)\big\}
997:     \ge \xi_n(k)\xi_n(l).
998:   \end{equation}
999:   Since $k+l\ge g(n)$ we can, in view of \eqref{e:g}, ignore the term
1000:   $\xi_n(k+l)$ on the left-hand side. Taking $s$ sufficiently large then
1001:   proves \eqref{e:submult}. This concludes the proof of the lemma.
1002: \end{proof}
1003: 
1004: We are now ready to complete the proof of Proposition~\ref{p:sum}.
1005: 
1006: \subsection{Proof of Proposition~\ref{p:sum}}
1007: \label{s:propproved}
1008: 
1009:   We shall establish that
1010:   \begin{equation}
1011:     \lim_{n\to\infty}
1012:     \max_{x\in \mathcal V_n}
1013:     \bigg|
1014:     \psum_{y_1,\dots,y_i\in B_n\setminus x}
1015:     \mathbb P_x\Big[\bigcap_{j=1}^i H_n(y_j)< a m(n)\Big]
1016:     -a^i\bigg| =0.
1017:   \end{equation}
1018:   Observe that, with $y_0=x$,
1019:   \begin{equation}
1020:     \begin{split}
1021:       \label{e:longsum}
1022:       \psum_{y_1,\dots,y_i\in B_n\setminus x}
1023:       \mathbb P_x\Big[&\bigcap_{j=1}^i H_n(y_j)< a m(n)\Big]
1024:       =
1025:       i!
1026:       \psum_{y_1,\dots,y_i\in B_n\setminus x}
1027:       \mathbb P_x\big[\bar H_n(y_1,\dots,y_i)< a m(n)\big]
1028:       \\&=
1029:       i! \sum_{d_1,\dots,d_i=1}^n
1030:       \psum_{\substack{y_1,\dots,y_i\in B_n\\d(y_i,y_{i-1})=d_i}}
1031:       \mathbb P_x\big[\bar H_n(y_1,\dots,y_i)< a m(n)\big].
1032:     \end{split}
1033:   \end{equation}
1034:   Consider first the summation over distances larger than $g(n)$. Using
1035:   Lemma~\ref{l:covergood} we get that (uniformly in the starting position $x$)
1036:   \begin{equation}
1037:     \begin{split}
1038:       i! \sum_{d_1,\dots,d_i=g(n)}^n&
1039:         \psum_{\substack{y_1,\dots,y_i\in B_n\\d(y_i,y_{i-1})=d_i}}
1040:         \mathbb P_x\big[\bar H_n(y_1,\dots,y_i)< a m(n)\big]
1041:         \\&=
1042:         \sum_{d_1,\dots,d_i=g(n)}^n
1043:         \psum_{\substack{y_1,\dots,y_i\in B_n\\d(y_i,y_{i-1})=d_i}}
1044:         2^{-in} m(n)^i a^i(1+o(1))
1045:         =a^i(1+o(1)).
1046:       \end{split}
1047:   \end{equation}
1048:   For the second equality we used the fact that by \eqref{e:Bsize},
1049:   \eqref{e:vcond} and the finiteness of $i$ there are
1050:   $2^n m(n)^{-1}(1+o(1))$ choices for every $y_i$.
1051: 
1052:   To estimate the remaining contribution to \eqref{e:longsum},
1053:   we first bound  the sum
1054:   \begin{equation}
1055:     \label{e:smalls}
1056:     \sum_{d=1}^{g(n)-1}
1057:     \sum_{y\in B_n:d(y,x)=d}
1058:     C e^a  \big(2^{-n}m(n) + \xi_n(d)\big)
1059:     \le
1060:     \sum_{d=1}^{g(n)-1} v_n(d)\xi_n(d)
1061:     +V_n(g(n)-1)  2^{-n}m(n).
1062:   \end{equation}
1063:   Both summands of in the last formula converge to $0$ which can be seen
1064:   easily from  \eqref{e:vcond} and  \eqref{e:Bsize}. For $d>g(n)$,
1065:   by Lemma~\ref{l:xi}(iii), $\xi_d(n)\ll 2^{-n}m(n)$.
1066:   Therefore, using \eqref{e:Bsize}, for all $n$ large enough
1067:   \begin{equation}
1068:     \label{e:bigs}
1069:     \sum_{d=1}^{n}
1070:     \sum_{y\in B_n:d(y,x)=d} C e^a\big(2^{-n}m(n) + \xi_n(d)\big)
1071:     \le 2Ce^a|B_n|2^{-n}m(n)\le 4 C e^a.
1072:   \end{equation}
1073:   According to  Lemma~\ref{l:coverbad},  the remaining
1074:   part of \eqref{e:longsum} then  satisfies
1075:   \begin{equation}
1076:     \begin{split}
1077:       i!& \sum_{\substack{d_1,\dots,d_i=1\\\exists d_i\le g(n)}}^n
1078:       \psum_{\substack{y_1,\dots,y_i\in B_n\\d(y_i,y_{i-1})=d_i}}
1079:       \mathbb P_x\big[\bar H_n(y_1,\dots,y_i)< a m(n)\big]
1080:       \\&\le
1081:       i!i
1082:       \sum_{d_1=1}^{g(n)}
1083:       \sum_{\substack{y\in B_n\\d(y,x)=d_1}}
1084:       C a \big(2^{-n}m(n)+ \xi_n(d_1)\big)
1085:       \bigg( \sum_{d=1}^{n}
1086:         \sum_{\substack{y\in B_n\\d(y,x)=d}}
1087:         C a\big(2^{-n}m(n)+ \xi_n(d)\big)
1088:         \bigg)^{i-1},
1089:       \end{split}
1090:   \end{equation}
1091:   which converges to $0$ by \eqref{e:smalls} and \eqref{e:bigs}. This
1092:   finishes the proof of Proposition~\ref{p:sum}.
1093: %%%% END OF ht2.tex %%%%
1094: 
1095: %%%% START OF ht3.tex %%%%
1096: \section{Proof of Theorem~\ref{t:perc} and of Theorem~\ref{t:sampling}}
1097: \label{s:equiv}
1098: 
1099: 
1100: In this section we apply Theorem~\ref{t:gen} to derive the
1101: asymptotic hitting distribution of randomly chosen sets in two
1102: different settings: for random clouds (namely we prove
1103: Theorem~\ref{t:perc}) and in the setting of drawing without
1104: replacement (which is Theorem~\ref{t:sampling}).
1105: 
1106: 
1107: \begin{proof}[Proof of  Theorem~\ref{t:perc}]
1108: To prove Theorem~\ref{t:perc} we will naturally show that the
1109: assumptions of
1110: Theorem~\ref{t:gen} are satisfied for percolation clouds $A_n$ of
1111: density $\bar m(n)^{-1}$, where $n\log n\ll \bar m(n)\ll 2^n(\log n)^{-1}$.
1112: 
1113: We first verify condition \eqref{e:Bsize}, i.e.~that
1114: $P$-a.s.~$|A_n|=2^n \bar m(n)^{-1}(1+o(1))$.
1115: By Chebyshev exponential inequality, for any $\delta >0$
1116: with $\lambda >0$,
1117: \begin{equation}
1118:   \begin{split}
1119:     \label{e:cheb}
1120:      P[|A_n|\gtrless(1\pm \delta )2^nn^{-1}]
1121:     &\le
1122:     \exp\{\mp\lambda  (1\pm \delta )2^n\bar m(n)^{-1}\}
1123:     (1+(e^{\pm \lambda }-1)\bar m(n)^{-1})^{2^n}
1124:     \\&\le
1125:     \exp\{\mp\lambda  (1\pm \delta )2^n\bar m(n)^{-1} +
1126:       (e^{\pm \lambda }-1)2^n\bar m(n)^{-1}\}.
1127:   \end{split}
1128: \end{equation}
1129: Taking $\lambda $ sufficiently small and using the fact that
1130: $2^n\bar m(n)^{-1}\gg \log n$, we see that  the right-hand side of the last
1131: equation is summable. Borel-Cantelli lemma then yields the result.
1132: 
1133: Let $f_n(k)=n(\log \bar m(n))^{-1}\bbone\{k=1\} + n\bbone\{k>1\}$. To prove
1134: that the first part of \eqref{e:vcond} is satisfied for the percolation
1135: cloud we show:
1136: \begin{lemma}
1137:   \label{l:cloudv}
1138:   There exists $C$ large enough, such that for a.e.~realisation of $A_n$
1139:   and for $n$ large enough
1140:   \begin{equation}
1141:     \label{e:rhs}
1142:     v_n(k)\le C\Big[\binom n k \bar m(n)^{-1} +  f_n(k)\Big]\qquad\forall
1143:     k\in\{1,\dots,g(n)\}.
1144:   \end{equation}
1145: \end{lemma}
1146: \begin{proof}
1147:   Let $F_n(k)$ denote the right-hand side of
1148:   \eqref{e:rhs}. By definition of $v_n(k)$,
1149:   \begin{equation}
1150:     \begin{split}
1151:       P[v_n(k)\ge F_n(k)]
1152:       &\le
1153:       \sum_{x\in \mathcal V_n}
1154:       P[|\{y\in A_n:d(x,y)=k\}|\ge F_n(k)]
1155:       \\&=
1156:       2^n
1157:       P[|\{y\in A_n:d(\boldsymbol 0,y)=k\}|\ge F_n(k)].
1158:     \end{split}
1159:   \end{equation}
1160:   Using the same calculation as in \eqref{e:cheb} this is bounded from
1161:   above by
1162:   \begin{equation}
1163:     2^n \exp\Big\{-C\lambda \Big[\binom n k \bar m(n)^{-1} + f_n(k)\Big]+
1164:       (e^\lambda -1)\binom n k \bar m(n)^{-1}\Big\}.
1165:   \end{equation}
1166:   If we choose $\lambda = \log \bar m(n)$ and $C$ large enough for $k=1$, or
1167:   $\lambda = \mathit{const}$ and $C$ large enough for $k>1$, then the
1168:   right-hand side of the last equation decays at least as fast as $e^{-cn}$ for all
1169:   $k$. Summing over $k$ and using Borel-Cantelli Lemma yields the desired
1170:   result.
1171: \end{proof}
1172: 
1173: Lemma~\ref{l:cloudv} implies that
1174: \begin{equation}
1175:   \sum_{k=1}^{g(n)-1} v_n(k) \xi_n(k)
1176:   \le
1177:   \sum_{k=1}^{g(n)-1}  C\Big[\binom n k \bar m(n)^{-1}+ f_n(k)\Big] \xi_n(k).
1178: \end{equation}
1179: Using Lemma~\ref{l:xi}(i),(iv) this can be bounded by
1180: \begin{equation}
1181:   \label{e:condcond}
1182:   C\bigg\{
1183:     \sum_{k=1}^{g(n)-1}
1184:     \xi_n(k)\binom n k \bar m(n)^{-1}+
1185:     \sum_{k=1}^3 f_n(k)\binom n k^{-1}
1186:     +\sum_{k=4}^{g(n)-1} n  f_n(k) \binom n k^{-1}
1187:     \bigg\}
1188: \end{equation}
1189: The last two terms in the last formula are bounded by
1190: \begin{equation}
1191:     C \sum_{k=1}^3 f_n(k) n^{-k}
1192:     + C n^3 n^{-4}\xrightarrow{n\to\infty}0
1193: \end{equation}
1194: as can be seen easily from the definition of $f_n(k)$.
1195: The first term in \eqref{e:condcond} equals
1196: \begin{equation}
1197:   \frac {n2^{-n}}{2\bar m(n)}
1198:   \sum_{k=1}^{g(n)-1}
1199:   \sum_{j=1}^{n-k}
1200:   \binom n { j+k} \frac 1j
1201:   =
1202:   \frac {n}{2\bar m(n)}
1203:   \sum_{j=1}^{n-1}
1204:   \frac {2^{-n}}j
1205:   \sum_{k=j+1}^{n\wedge (g(n)+j-1)}\binom n k\le
1206:   \frac {C n \log n }{\bar m(n)}.
1207: \end{equation}
1208: This tends to $0$ by the assumptions on $\bar m(n)$.
1209: Therefore $v_n(k)$ verifies the first part of \eqref{e:vcond} $P$-a.s.~.
1210: 
1211: To verify the second part observe first that if
1212: $\bar m(n)\ge \delta 2^n n^{-1}$ for some $\delta >0$, then \eqref{e:gg}
1213: holds for $g(n)=2$.  Therefore,
1214: by Lemma~\ref{l:cloudv}, $V_n(g(n)-1)\le v_n(1) + 1 \le C \ll |A_n|$.
1215: We can hence
1216: further suppose that $\bar m(n)\ll 2^n n^{-1}$.
1217: By moderate deviations argument, and since $n/2-g(n)\gg n^{1/2}$,
1218: \begin{equation}
1219:   |\{y:d(\boldsymbol 0,y)\le g(n)-1\}|\le 2^n e^{-c g(n)^2/n}.
1220: \end{equation}
1221: Since $2^n\bar m(n)^{-1}\gg n$ there is a function $f(n)$ such that
1222: \begin{equation}
1223:   2^n \bar m(n)^{-1}f(n)\gg n
1224:   \qquad \text{and}\qquad
1225:   1\gg f(n)\gg e^{-cg(n)^2/n}.
1226: \end{equation}
1227: As in \eqref{e:cheb}
1228: \begin{equation}
1229:   P\Big[V_n(g(n)-1)\ge
1230:     \frac{2^{n}f(n)}{\bar m(n)}\Big]\le
1231:   2^n\exp\Big\{\frac{2^n}{\bar m(n)}\big[
1232:     -\lambda f(n)+
1233:     (e^\lambda -1)e^{-cg(n)^2/n}
1234:     \big]
1235:     \Big\}.
1236: \end{equation}
1237: For our choice of $f$ this is summable. Therefore a.s.~for $n$ large enough
1238: $V_n(g(n)-1)\le 2^n \bar m(n)^{-1} f(n)\ll |A_n|$. This verifies the second part of
1239: \eqref{e:vcond}.
1240: 
1241: We have verified that with $P$ probability one the sequence of percolation
1242: clouds $A_n$ satisfies all the assumptions of Theorem~\ref{t:gen}.
1243: This proves Theorem~\ref{t:perc}.
1244: \end{proof}
1245: 
1246: \begin{proof}[Proof of  Theorem~\ref{t:sampling}]
1247:   For any $m(n)$ satisfying the conditions of Theorem~\ref{t:sampling} it
1248:   is possible to choose $\bar m(n)$ satisfying the conditions of
1249:   Theorem~\ref{t:perc} such that
1250:   \begin{equation}
1251:     \bar m(n)^{-1} \ge (1+\varepsilon )m(n).
1252:   \end{equation}
1253:   We now consider a sequence of percolation clouds $A_n$ with density
1254:   $\bar m(n)^{-1}$ defined on the same probability space
1255:   $(\Omega',F',P')$ as $A'_n$. Since $P'$-a.s.~for all $n$ large enough
1256:   $|A_n|>2^n m(n)^{-1}=|A'_n|$, we can couple $A'_n$ and $A_n$ in the way
1257:   that for all $n$ large $A'_n\subset A_n$. Moreover, $A_n$ satisfies the
1258:   conditions \eqref{e:volumes}--\eqref{e:vcond} of Theorem~\ref{t:gen}. To
1259:   finish the proof observe that if $A_n$ satisfies these conditions, then
1260:   any subset of $A_n$ satisfies them too.
1261: \end{proof}
1262: %%%% END OF ht3.tex %%%%
1263: 
1264: %\input htapp
1265: 
1266: %%%% START OF htbiblio.tex %%%%
1267: \subsection*{Acknowledgements} Both authors thank
1268: the Chair of Stochastic Modelling
1269: of the \'Ecole Polytechnique F\'ed\'erale of Lausanne
1270: for financial support. Ji\v r\'\i~\v Cern\'y
1271: thank the Centre de Physique Th\'eorique of Marseille for hospitality.
1272: 
1273: 
1274: \bibliographystyle{amsalpha}
1275: \bibliography{ht0}
1276: %%%% END OF htbiblio.tex %%%%
1277: 
1278: 
1279: 
1280: \end{document}
1281: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1282: % vim:foldmethod=marker:foldnestmax=1:foldlevel=0:foldmarker=%<<<,%>>>
1283: