1: \documentclass[10pt]{article}
2:
3: \usepackage{amsfonts}
4: \usepackage{amsxtra}
5: \usepackage{amsmath}
6: \usepackage{a4}
7: \usepackage{amssymb}
8: \usepackage{theorem}
9: \usepackage{epsfig}
10: \oddsidemargin=10mm
11: \textwidth=142mm
12: \topmargin=0mm
13: \textheight=220mm
14:
15: \renewcommand{\baselinestretch}{1.2}
16:
17: \newtheorem{theorem}{Theorem}
18: \newtheorem{lemma}[theorem]{Lemma}
19: \newtheorem{corollary}[theorem]{Corollary}
20: {\theorembodyfont{\rmfamily}\newtheorem{example}[theorem]{Example}}
21: {\theorembodyfont{\rmfamily}\newtheorem{remark}[theorem]{Remark}}
22: \newtheorem{proposition}[theorem]{Proposition}
23: \numberwithin{theorem}{section}
24: \numberwithin{equation}{section}
25: \numberwithin{figure}{section}
26:
27: \newcommand{\given}{\big\vert}
28:
29: \newcommand{\Bin}{\text{Bin}}
30: \newcommand{\Ber}{\text{Ber}}
31:
32: \newcommand{\intersect}{\cap}
33:
34: \newcommand{\bu}{\mathbf{u}}
35: \newcommand{\by}{\mathbf{y}}
36: \newcommand{\bv}{\mathbf{v}}
37:
38: \newcommand{\RR}{\mathbb{R}}
39: \newcommand{\PP}{\mathbb{P}}
40: \newcommand{\ZZ}{\mathbb{Z}}
41: \newcommand{\NN}{\mathbb{N}}
42: \newcommand{\EE}{\mathbb{E}\,}
43: \newcommand{\CC}{\mathbb{C}}
44:
45: \newcommand{\cC}{\mathcal{C}}
46: \newcommand{\cL}{\mathcal{L}}
47: \newcommand{\cB}{\mathcal{B}}
48: \newcommand{\cS}{\mathcal{S}}
49: \newcommand{\cT}{\mathcal{T}}
50:
51: \renewcommand{\Box}{\square}
52: \newcommand{\ty}{{\tilde{y}}}
53:
54:
55: \newcommand{\tC}{{\tilde{C}}}
56: \newcommand{\tu}{{\tilde{u}}}
57: \newcommand{\tA}{{\tilde{A}}}
58: \newcommand{\tT}{{\tilde{T}}}
59: \newcommand{\tM}{{\tilde{M}}}
60: \newcommand{\tY}{{\tilde{Y}}}
61: \newcommand{\tV}{{\tilde{V}}}
62: \newcommand{\tS}{{\tilde{S}}}
63: \newcommand{\tX}{{\tilde{X}}}
64:
65: \newcommand{\Ubar}{{\bar{U}}}
66:
67: \newcommand{\bone}{\boldsymbol{1}}
68:
69: \newcommand{\toi}{\rightarrow \infty}
70: \newcommand{\psis}{F_{k_{1},\dots,k_{n}}}
71: \newcommand{\bfx}{{\bf x}}
72: \newcommand{\ep}{{\epsilon}}
73: \newcommand{\bsl}{\backslash}
74: \newcommand{\bx}{\mathbf{x}}
75: \newcommand{\bfi}{{\bf i}}
76: \newcommand{\bfj}{{\bf j}}
77: \newcommand{\bfu}{{\bf u}}
78: \newcommand{\bfv}{{\bf v}}
79: \newcommand{\bl}{{\bf L}}
80: \newcommand{\bpi}{{\bf \pi}}
81: \newcommand{\ba}{{\bf a}}
82: \newcommand{\adn}{a_{\Delta_{\nu}}}
83: \newcommand{\be}{\mbox{$\mathbb{E}$}}
84: \newcommand{\bp}{\mbox{$\mathbb{P}$}}
85: \newcommand{\ce}{\mbox{$\cal E$}}
86: \newcommand{\cf}{\mbox{$\cal F$}}
87: \newcommand{\cdd}{\mbox{${\cal D}_0$}}
88: \newcommand{\tce}{\tilde {\mbox{$\cal E$}}}
89: \newcommand{\tcf}{\tilde {\mbox{$\cal F$}}}
90: \newcommand{\hce}{\hat {\mbox{$\cal E$}}}
91: \newcommand{\hcf}{\hat {\mbox{$\cal F$}}}
92: \newcommand{\tm}{\tilde{m}}
93: \newcommand{\taf}[1][\partial F]{\tau_{#1}}
94: \newcommand{\tF}{\tilde{F}}
95: \newcommand{\tB}{\tilde{B}}
96: \newcommand{\tphi}{\tilde{\phi}}
97: \newcommand{\tE}{\tilde{E}}
98: \newcommand{\hf}{{\hat F}}
99: \newcommand{\hc}{{\hat{c}}}
100: \newcommand{\hmu}{\hat {\mu}}
101: \newcommand{\cd}{\mbox{$\cal D$}}
102: \newcommand{\cc}{\mbox{$\cal C$}}
103: \newcommand{\hcfi}{\mbox{$\hat{\cal F}_{\infty}$}}
104: \newcommand{\che}{\mbox{$\hat{\cal E}$}}
105: \newcommand{\chf}{\mbox{$\hat{\cal F}$}}
106: \newcommand{\bz}{\mbox{$\mathbb Z$}}
107: \newcommand{\br}{\mbox{$\mathbb R$}}
108: \newcommand{\bn}{\mbox{$\mathbb N$}}
109: \newcommand{\bpt}{\bar{p}_{t}}
110: \newcommand{\hK}{\hat K}
111: \newcommand{\gi}{\Gamma^{(i)}}
112: \newcommand{\gj}{\Gamma^{(j)}}
113: \newcommand{\mi}{\mu^{(i)}}
114: \newcommand{\mj}{\mu^{(j)}}
115: \newcommand{\cei}{\mbox{${\cal E}^{(i)}$}}
116: \newcommand{\cej}{\mbox{${\cal E}^{(j)}$}}
117: \newcommand{\cfi}{\mbox{${\cal F}^{(i)}$}}
118: \newcommand{\cfio}{\mbox{${\cal F}_1^{(i)}$}}
119: \newcommand{\cfj}{\mbox{${\cal F}^{(j)}$}}
120: \newcommand{\dwi}{d_w^{(i)}}
121: \newcommand{\pe}{P^x_{\epsilon}}
122: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
123:
124: \makeatletter
125: \@addtoreset{equation}{section}
126: \renewcommand{\theequation}{\thesection.\arabic{equation}}
127: \makeatother
128:
129: \def\enpf{
130: { \parfillskip=0pt\hfil {\hbox{$\mbox{$\Box$}$}} \par\bigskip }
131: }
132:
133: \allowdisplaybreaks
134:
135: \begin{document}
136:
137: \title{Heavy tails in last-passage percolation}
138: \author{
139: \textbf{$\,\,\,\,\,\,\,\,$Ben Hambly}
140: \hspace{1cm}
141: \textbf{James B.~Martin}\\
142: \textit{University of Oxford}
143: }
144: \date{March 21, 2006}
145: \maketitle
146:
147: \renewcommand{\thefootnote}{}
148: \footnotetext[1]{
149: $\!\!\!\!\!$
150: \textit{MSC 2000 subject classifications:} Primary 60K35; Secondary 82B41.
151: \textit{Keywords and phrases:}
152: Last-passage percolation, heavy tails, Airy process,
153: regular variation, stable process, multifractal spectrum.}
154:
155: \begin{abstract}
156: We consider last-passage percolation models in two dimensions,
157: in which the underlying weight distribution has
158: a heavy tail of index $\alpha<2$.
159: We prove scaling laws and asymptotic distributions,
160: both for the passage times and for the shape of
161: optimal paths; these are expressed in
162: terms of a family (indexed by $\alpha$) of
163: ``continuous last-passage percolation'' models in the unit square.
164: In the extreme case $\alpha=0$ (corresponding to
165: a distribution with slowly varying tail)
166: the asymptotic distribution of the optimal path
167: can be represented by a random self-similar measure on $[0,1]$,
168: whose multifractal spectrum we compute.
169: By extending the continuous last-passage percolation
170: model to $\RR^2$ we obtain
171: a heavy-tailed analogue of the Airy process,
172: representing the limit of appropriately scaled vectors
173: of passage times to different points in the plane.
174: We give corresponding results for a directed
175: percolation problem based on $\alpha$-stable L\'evy processes,
176: and indicate extensions of the results to higher dimensions.
177: \end{abstract}
178:
179: \section{Introduction}
180: Directed last-passage percolation in two dimensions has received much
181: attention in recent years. In certain specific cases, for example
182: where the weights at each site are i.i.d.\ with exponential or
183: geometric distribution, very precise scaling laws and asymptotic
184: distributions are now known, both for the passage times and for the
185: shape of optimal paths (see for example \cite{Johshape,
186: Johanssontransversal, BDMMZ}). Such cases are closely related to the
187: longest increasing subsequence problem, and to Markovian interacting
188: particle systems such as the totally asymmetric exclusion process;
189: there are also very close links to random matrix theory, for example
190: to the behaviour of the largest eigenvalue of a large matrix drawn
191: from the Gaussian Unitary Ensemble (see for example
192: \cite{Neilsurvey} for a survey).
193:
194: It is believed that the behaviour proved for the exponential and
195: geometric cases should be \textit{universal}, in that the same scaling
196: laws and asymptotic distributions should occur in the last-passage
197: percolation model whose underlying weight distribution is from a much
198: more general class (certainly including any distribution with an
199: exponentially decaying tail, and maybe also those with sufficiently
200: light polynomial tails). The growth models corresponding to these
201: last-passage percolation problems should belong to the
202: \textit{Kardar-Parisi-Zhang (KPZ) universality class} (see for example
203: \cite{KrugSpohn}). However, only very limited universality results
204: have been proved: for example, conditions under which laws of large
205: numbers for the passage times (or ``shape theorems'') hold, and
206: asymptotics for passage-times close to the boundary of the quadrant
207: \cite{JBMshape, BodineauMartin, BaiSui}.
208:
209: In this paper, we study cases in which the tail of the weight
210: distribution is sufficiently heavy that such shape theorems fail, and
211: which certainly fall outside the universality class described
212: above. Specifically, we assume that the tail of the weight
213: distribution is regularly varying with index $\alpha<2$. We describe
214: a family of ``continuous last-passage percolation'' models (indexed by
215: $\alpha$), and use them to provide scaling laws and asymptotic
216: distributions for the discrete models, both for the passage times and
217: for the shape of the optimal paths. Thus we have a universality
218: result for these heavy-tailed models as the only information
219: required to determine the scaling limits is the parameter $\alpha$.
220:
221: One example of an application where such a heavy-tailed assumption is
222: very natural is in the use of last-passage models to represent
223: networks of \textit{queues in tandem}. The vertex weights in the
224: percolation models correspond to service times in the queueing
225: systems, and passage times in the percolation models correspond
226: to the total time spent in the queueing system by particular customers;
227: see for example \cite{GlyWhi, BBM, blocking}.
228:
229: In Section \ref{main}, we define the discrete last-passage
230: percolation model
231: precisely; we then describe the continuous last-passage
232: model and state our main convergence results. We also derive from the
233: continuous model a stationary process which can be seen as a
234: heavy-tailed analogue of the \textit{Airy process} (which was
235: developed by Pr\"ahofer and Spohn \cite{PraSpo} and Johansson
236: \cite{JohanssonAiry}), and which gives a process limit for vectors of
237: passage times to different points, appropriately scaled.
238:
239: The proofs of the convergence results for passage-times are given in
240: Section \ref{Tsection}, and those for the optimal paths are given in
241: Section \ref{pathsection}. The results on the heavy-tailed Airy
242: process are proved in section \ref{Airysection}.
243:
244: In Section \ref{greedysection} we explore the case where the tail of
245: the weight distribution is \textit{slowly varying} (i.e.\ $\alpha=0$).
246: It is no longer possible to provide a non-degenerate limiting
247: distribution for the passage times; however, asymptotics for the form
248: of the optimal paths are still possible, and in fact the
249: distribution of the limiting path that
250: arises can be described in a particularly simple and
251: algorithmic way. As the path is increasing it can be thought of
252: as the distribution of a random measure on $[0,1]$;
253: this measure is self-similar and we compute its multifractal
254: spectrum.
255:
256: The \textit{Brownian directed percolation} model has recently been
257: much studied in various contexts (see for example \cite{GUEs,
258: GraTraWid, hmo, OcoYor}). In Section \ref{stablesection} we discuss a
259: related model in which Brownian motion is replaced by an
260: $\alpha$-stable L\'evy process, and we again prove distributional
261: convergence to the continuous last passage percolation model.
262:
263: The bulk of the paper concerns the case of two-dimensional
264: last-passage percolation. However, almost all of the results extend
265: easily to $d$ dimensions, $d\geq 3$, and now apply for $\alpha<d$. We
266: indicate these extensions in Section \ref{dsection}.
267:
268: Simulations of trees of optimal paths, of the limiting path for
269: $\alpha=0$, and of heavy-tailed Airy processes are given in Sections
270: \ref{main}, \ref{greedysection} and \ref{Airysection}.
271:
272: \section{Main results}\label{main}
273: \subsection{Definition of the discrete problem}
274:
275: Let $F$ be a distribution function.
276: We will assume that the tail of the distribution $F$
277: is \textit{regularly varying} with index $\alpha\in(0,2)$; that is,
278: for all $t>0$,
279: \[
280: \frac{1-F(tx)}{1-F(x)}\to t^{-\alpha} \text{ as } x\to\infty.
281: \]
282: We will also assume (merely for convenience)
283: that $F$ is a continuous distribution and $F(0)=0$.
284:
285: The discrete last-passage percolation model with
286: underlying weight distribution $F$ is usually defined as follows.
287:
288: Let $X(i,j), i,j\in\NN$ be i.i.d.\ with common distribution $F$.
289: The quantity $X(i,j)$
290: represents the \textit{weight} at the site $(i,j)\in\ZZ_+^2$.
291:
292: For $n\in\NN$, we will define the quantity $T^{(n)}$, the
293: \textit {last-passage time}
294: between $(1,1)$ and $(n,n)$. Let $\Pi_n$ be the set of
295: \textit{directed paths} between $(1,1)$ and $(n,n)$. Each such path
296: begins at $(1,1)$ and ends at $(n,n)$, and each step consists of
297: increasing one of the two coordinates by 1.
298: That is, for any $\pi\in\Pi_n$, we can write $\pi=(v_1, v_2, \dots,
299: v_{2n})$, where $v_1=(1,1)$, $v_{2n}=(n,n)$, and, for each
300: $i=1,\dots,2n-1$, $v_{i+1}-v_i$ is either $(1,0)$ or $(0,1)$.
301:
302: The weight of such a path is the sum of the weights $X(i,j)$
303: associated with the points $(i,j)$ in the path. Then $T^{(n)}$ is the
304: maximal weight of a directed path between $(1,1)$ and $(n,n)$; that is:
305: \begin{equation}
306: \label{Tndef1}
307: T^{(n)}=\max_{\pi\in\Pi_n} \sum_{v\in\pi} X(v).
308: \end{equation}
309: Note that $T^{(n)}$ depends only on the weights $X(v)$,
310: $v\in\{1,\dots,n\}^2$.
311:
312: \subsection{Continuous model}
313:
314: We start with an alternative representation of the discrete model.
315: Let $M^{(n)}_1 \geq M^{(n)}_2 \geq \dots \geq M^{(n)}_{n^2}$ be the
316: order statistics, written in decreasing order, from an i.i.d.\ sample
317: of size $n^2$ from the distribution $F$.
318:
319: Consider the set $\{1/n, 2/n, \dots, (n-1)/n, 1\}^2 \subset [0,1]^2$,
320: of size $n^2$. Let the sequence $Y^{(n)}_1, \dots, Y^{(n)}_{n^2}$
321: consist of a random ordering of the points of this set, chosen
322: uniformly from the $(n^2)!$ possibilities.
323: We regard $Y^{(n)}_i$ as the location of the $i$th largest weight
324: $M^{(n)}_i$ (and we have scaled so that all points lie in the box
325: $[0,1]^2$).
326:
327: For two points $y, y'\in[0,1]^2$, we say that $y$ and $y'$ are
328: \textit{compatible}, and write $y\sim y'$, if $y,y'$ are partially
329: ordered in that \textit{either} $y\leq
330: y'$ co-ordinatewise, \textit{or} $y'\leq y$
331: co-ordinatewise. (Informally, one of $y$ and $y'$ is below and to the
332: left of the other).
333: An increasing path will consist of a set of points such that every
334: pair of points in the set is compatible. We describe the collection
335: of increasing paths by the collection $\cC^{(n)}$, which depends on
336: the points $Y^{(n)}_1, \dots, Y^{(n)}_{n^2}$ alone:
337: \begin{equation}
338: \label{Cndef}
339: \cC^{(n)}=
340: \cC^{(n)}(Y^{(n)}_1,\dots, Y^{(n)}_{n^2})=
341: \left\{A\subseteq\{1,\dots,n^2\}
342: \text{ such that for all } i,j\in A,
343: Y^{(n)}_i\sim Y^{(n)}_j
344: \right\}.
345: \end{equation}
346:
347: Now we can give a new definition for $T^{(n)}$, equivalent (in
348: distribution) to (\ref{Tndef1}):
349: \begin{equation}\label{Tndef2}
350: T^{(n)}=\max_{A\in\cC^{(n)}}\sum_{i\in A} M^{(n)}_i.
351: \end{equation}
352:
353: We formulate the limiting continuous model by defining the
354: distribution of a random variable $T$ in an analogous way.
355:
356: First let $Y_1, Y_2, \ldots$ be an i.i.d.\ sequence, with each $Y_i$
357: uniformly distributed on the square $[0,1]^2$.
358: Let $W_1, W_2, \ldots$ be an i.i.d.\ sequence of exponential random
359: variables with mean 1 (independent of the $(Y_i)$). Now write, for
360: each $k\in\NN$, $M_k=(W_1+\ldots+W_k)^{-1/\alpha}$. (Then with
361: probability 1, $M_k>M_{k+1}$ for each $k$, and $M_k\to0$ as
362: $k\to\infty$). The motivation for this definition
363: is given by equation (\ref{Mconv}) below.
364: $M_k$ is the $k$th largest weight, which we imagine positioned
365: at the point $Y_k\in[0,1]^2$. (The set of locations $Y_k$
366: is of course dense in $[0,1]^2$ with probability 1).
367:
368: Analogously to (\ref{Cndef}), we
369: represent the set of increasing paths by the collection $\cC$:
370: \begin{equation}\label{cCdef}
371: \cC=
372: \cC(Y_1, Y_2, \dots)=
373: \left\{A\subseteq\{1,2,\dots\}
374: \text{ such that for all } i,j\in A,
375: Y_i\sim Y_j
376: \right\}.
377: \end{equation}
378: Then define
379: \begin{equation}\label{Tdef}
380: T=\sup_{A\in\cC}\sum_{i\in A} M_i.
381: \end{equation}
382:
383: \noindent\textit{Remark:}
384: Note that one could equivalently define $T$ in (\ref{Tdef})
385: as the sup of the weight of \textit{finite} increasing paths $A$,
386: since either the sup is finite in which case
387: the weight of any infinite path can be arbitrarily closely
388: approximated by that of a finite path, or the sup is infinite
389: in which case one can find a finite path with an arbitrarily large
390: weight. In particular
391: $T$ can be seen as the supremum of a countable family of measurable
392: random variables, and so is itself measurable.
393: In Section \ref{Airysection} an equivalent construction
394: of the continuous last-passage problem
395: is given using a Poisson random measure approach
396: rather than the sequence of i.i.d.\ uniform positions
397: in the unit square described above.
398:
399: \subsection{Convergence results}
400: First note that for all $k$,
401: \[ \left( Y^{(n)}_1, Y^{(n)}_2,\dots Y^{(n)}_k \right) \to
402: \left(Y_1, Y_2,\dots,Y_k\right) \]
403: in distribution, as $n\to\infty$.
404:
405: Now define $a_N=F^{(-1)}\left(1-\frac{1}{N}\right)$.
406: (As an example,
407: if the weight distribution $F$ is Pareto($\alpha$), with
408: $F(x)=1-x^{-\alpha}$, then $a_{N}=N^{1/\alpha}$. In general,
409: $\lim_{N\to\infty} \log a_{N}/\log N=1/\alpha$).
410:
411: Recall that $M^{(n)}_k$ is
412: the $k$th largest value from a sample of size $n^2$
413: from the distribution $F$.
414: We write $\tM^{(n)}_i=a_{n^2}^{-1} M^{(n)}_i$.
415: Then from classical extreme value theory we have that, for all $k$,
416: \begin{equation}\label{Mconv}
417: \left( \tM^{(n)}_1, \tM^{(n)}_2,\dots \tM^{(n)}_k \right)
418: \to \left(M_1, M_2,\dots,M_k\right)
419: \end{equation}
420: in distribution, as $n\to\infty$
421: (see for example Section 9.4 of \cite{Davidbook}).
422:
423: In particular, $M^{(n)}_i$ is asymptotically of the order of
424: $a_{n^2}$, for any $i$.
425: (For example, for the Pareto distribution $F(x)=1-x^{-\alpha}$
426: mentioned above, we have
427: $a_{n^2}=n^{2/\alpha}$).
428: Since certainly $T^{(n)}\geq M^{(n)}_1$, we
429: have that $T^{(n)}$ grows asymptotically at least on the order of
430: $a_{n^2}$. In fact, we will show that this lower bound gives the right
431: order of magnitude.
432:
433: Specifically, let $\tT^{(n)}=a_{n^2}^{-1}T^{(n)}
434: =\sup_{A\in\cC^{(n)}}\sum_{i\in A} \tM^{(n)}_i$.
435: Then we will show:
436: \begin{theorem}
437: \label{Ttheorem}
438: The random variable $T$ defined at (\ref{Tdef})
439: is almost surely finite, and
440: $\tT^{(n)}\to T$ in distribution as $n\to\infty$.
441: \end{theorem}
442:
443: For comparison, one can consider the case of a lighter tail.
444: If $\int_0^\infty [1-F(x)]^{1/2}dx<\infty$
445: (this condition is very slightly stronger than the
446: existence of a finite second moment) then
447: a law of large numbers holds:
448: $n^{-1}T^{(n)}\to\gamma$ as $n\to\infty$
449: for some deterministic $\gamma$ \cite{animals}.
450: If the weights are exponential with mean 1,
451: then $\gamma=4$ \cite{Rost}, and then in fact
452: one has the much finer convergence result that
453: $n^{-1/3}(T^{(n)}-4n)$ converges in distribution
454: as $n\to\infty$, to the \textit{GUE Tracy-Widom distribution}
455: \cite{Johshape}.
456:
457:
458: \begin{figure}[ht]
459: \centering
460: \epsfig{figure=pareto1.eps, width=1.25\linewidth}
461: \caption{A simulation of the last-passage percolation
462: model with $F$ given by a Pareto distribution with index $1$.
463: The ``tree'' consisting of optimal paths from $(1,1)$
464: to $(i,j)$, for all $1\leq i,j\leq 159$ is displayed;
465: the thickened path is the optimal path from $(1,1)$
466: to $(159, 159)$. This represents $P^{(n)*}$ in the language
467: of Theorem \ref{paththeorem2}.
468: \label{paretofig}}
469: \end{figure}
470: \begin{figure}[ht]
471: \epsfig{figure=exponential.eps, width=1.25\linewidth}
472: \caption{As Figure \ref{paretofig}, but now with $F$
473: given by an exponential distribution. As $n$ grows,
474: the distribution of the optimal path from $(1,1)$ to $(n,n)$
475: becomes concentrated around the diagonal -- the deviations
476: of the path from a straight line are on the order of $n^{2/3}$.
477: This contrasts with the deviations on the order of $n$
478: observed in the heavy-tailed case, where the limiting
479: path distribution is non-degenerate (as illustrated in Figure
480: \ref{paretofig}).
481: \label{exponentialfig}}
482: \end{figure}
483:
484:
485: We now outline the results on
486: path convergence, which are given in full in Section
487: \ref{pathsection}.
488: We will show that the optimal
489: path for the continuous model is well defined;
490: that is, with probability 1 there exists a unique $A^*\in\cC$
491: such that $T=\sum_{i\in A^*} M_i$
492: (attaining the $\sup$ in (\ref{Tdef})).
493: Then there is a unique closed connected set $P^*\subset[0,1]^2$
494: which contains all the points $Y_i, i\in A^*$
495: and which itself has the directed path property
496: (i.e.\ $y\sim y'$ for all $y, y'\in P^*$).
497:
498:
499:
500: Analogously, let $A^{(n)*}$ be the optimal
501: path for the discrete model, achieving
502: the $\max$ in (\ref{Tndef2})
503: (Since the weight distribution $F$
504: is continuous, the finitely many increasing paths
505: all have different weights a.s., and
506: so this maximizing path is a.s.\ unique).
507: Let $P^{(n)*}$ be obtained by linear
508: interpolation between the locations
509: $\{Y_i^{(n)}: i\in A^{(n)*}\}$
510: of weights used in this optimal path, taken in increasing order.
511: Then we show that the distribution of $P^{(n)*}$ converges
512: to that of $P^*$ as $n\to\infty$ (under the Hausdorff metric
513: on closed subsets of $[0,1]^2$).
514:
515:
516:
517: In the corresponding situation for weights with
518: exponential distribution, the optimal path converges instead
519: to a trivial limit, the straight line from $(0,0)$ to $(1,1)$.
520: In general, the deviations of the optimal path from
521: $(1,1)$ to $(n,n)$ in the discrete model
522: are expected to be of the order of $n^{2/3}$ in cases
523: falling into the KPZ universality class
524: (proved rigorously in certain cases \cite{Johanssontransversal, BDMMZ}),
525: rather than on the order of $n$ as we see in the heavy-tailed case.
526: See Figures \ref{paretofig} and \ref{exponentialfig}
527: for simulations of optimal paths in the cases
528: of weight distributions which are Pareto and exponential.
529:
530: \subsection{Last-passage random fields and the heavy-tailed Airy process}
531: \label{preAiry}
532: In Section \ref{Airysection}
533: we show how the results described above
534: can be extended to give the multivariate convergence
535: of vectors of passage times to different points.
536:
537: For $x, y>0$, let
538: $T^{(n)}(x,y)$ be the maximal weight of a path from $(1,1)$ to
539: $(\lceil nx\rceil, \lceil ny\rceil)$.
540: (So for example the quantity $T^{(n)}$ defined in
541: (\ref{Tndef2}) is equal to $T^{(n)}(1,1)$).
542:
543: Define also $\tT^{(n)}(x,y)=a_{n^2}^{-1}T^{(n)}(x,y)$ as before.
544:
545: Then we will construct a random field $\{T(x,y), x, y>0\}$, using
546: a Poisson random measure construction rather than the sequence of
547: points ordered in decreasing order of weight above, in such a way that
548: \[ \Big\{\tT^{(n)}(x,y), x, y>0\Big\} \to \Big\{T(x,y),
549: x, y>0\Big\} \]
550: as $n\to\infty$, in the sense of convergence of finite-dimensional
551: distributions.
552:
553: From a scaling property of the distribution of the weights
554: one has further that the random field defined by
555: \[ \Theta(u,v)=\exp\left(-\frac{u+v}{\alpha}\right)
556: T\left(e^u, e^v\right) \]
557: is \textit{stationary} on $\RR^2$.
558: The convergence above can be rewritten as
559: \begin{equation}
560: \label{Thetaconv}
561: \Bigg\{ \exp\left(-\frac{u+v}{\alpha}\right)
562: \tT^{(n)}\left(e^u,e^v\right), u,v\in\RR\Bigg\}
563: \to \Big\{\Theta(u,v), u, v\in\RR\Big\}.
564: \end{equation}
565:
566: To remove the multiplicative factor
567: on the LHS of (\ref{Thetaconv}), one can look at a line
568: $u+v=\text{const}$; for example, the process $H_y=\Theta(y,-y)$.
569: This process is stationary in $y\in\RR$, and we obtain the
570: weak convergence
571: \[ \left\{ a_{n^2}^{-1} T^{(n)}\left(e^y,e^{-y}\right), y\in\RR\right\}
572: \to
573: \big\{\Theta(y,-y), y\in\RR \big\}.
574: \]
575: This gives an analogy with the ``Airy process'' \cite{JohanssonAiry}
576: \cite{PraSpo}, which arises for example in the case where the
577: underlying weight
578: distribution is exponential (with mean 1, say).
579: There one obtains
580: a stationary process limit
581: for the quantities
582: \[
583: \Big\{n^{-1/3} \Big[ T^{(n)}(1+yn^{-1/3}, 1-yn^{-1/3}) - 4n \Big],
584: \, y\in\RR \Big\},
585: \]
586: whose marginals are given by the GUE Tracy-Widom distribution.
587: Simulations of the ``heavy-tailed Airy process'' $H_y$
588: are given in Figures \ref{Airyfig1}-\ref{Airyfig3}.
589:
590: We also obtain estimates on the moments and correlations
591: of the random field $T$, showing for example that
592: $\EE T(x,y)^\beta<\infty$ for all $\beta<\alpha$
593: and giving bounds for $\EE|T(x,y)-T(x',y')|^\beta$.
594:
595: We note that the path convergence described above could
596: also be extended to the multivariate setting,
597: to describe the convergence of the distribution of
598: trees of optimal paths to a continuous tree structure
599: given by the set of optimal paths in the continuous
600: last-passage percolation model. However we do not
601: pursue this further in this paper.
602:
603: \section{Convergence of the last-passage
604: time distribution}
605: \label{Tsection}
606:
607: To establish the convergence in Theorem \ref{Ttheorem},
608: we will work with approximations to $T$ and $T^{(n)}$
609: which depend only on the $k$ largest weights.
610: First, define
611: \begin{gather*}
612: \cC_k= \cC(Y_1, Y_2, \dots, Y_k)= \left\{A\subseteq\{1,2,\dots,k\}
613: \text{ such that for all } i,j\in A, Y_i\sim Y_j \right\}. \\
614: \cC^{(n)}_k= \cC^{(n)}_k(Y^{(n)}_1,\dots, Y^{(n)}_{k\wedge n^2})=
615: \left\{A\subseteq\{1,\dots,k\wedge n^2\}
616: \text{ such that for all } i,j\in A,
617: Y^{(n)}_i\sim Y^{(n)}_j
618: \right\}.
619: \end{gather*}
620:
621: Note that in fact
622: $\cC_k
623: =\{A\in\cC: A\subseteq\{1,\dots,k\}\}
624: =\{A\cap\{1,\dots,k\}: A\in\cC\}$, and similarly for $\cC_k^{(n)}$.
625:
626: Now let
627: \[
628: T_k=\sup_{A\in\cC}\sum_{i\in A, i\leq k} M_i,
629: \]
630: and
631: \[
632: T^{(n)}_k=\sup_{A\in\cC^{(n)}}\sum_{i\in A, i\leq k} M^{(n)}_i.
633: \]
634: Note that indeed $T_k$ depends only on $(M_1,\dots, M_k)$
635: and $(Y_1,\dots, Y_k)$, while
636: $T^{(n)}_k$ depends only on $(M^{(n)}_1, \dots, M^{(n)}_k)$
637: and $(Y^{(n)}_1, \dots, Y^{(n)}_k)$.
638:
639: As before, define
640: $\tT^{(n)}_k=a_{n^2}^{-1}T^{(n)}_k$.
641:
642: We also define the ``remainder terms'' $S_k$, $S^{(n)}_k$ by
643: \[
644: S_k=\sup_{A\in\cC}\sum_{i\in A, i>k} M_i
645: \,\,
646: \text{ and }
647: \,\,
648: S^{(n)}_k=\sup_{A\in\cC^{(n)}}\sum_{i\in A, i>k} M^{(n)}_i.
649: \]
650: Write also $\tS^{(n)}_k=a_{n^2}^{-1}S^{(n)}_k$.
651:
652: \begin{lemma}
653: \label{Slemma}
654: With probability 1,
655: $S_k<\infty$ for all
656: $k\geq 0$, and $S_k\to 0$ as $k\to\infty$.
657: \end{lemma}
658:
659: In particular, putting $k=0$ we will have that $T<\infty$ a.s.
660: (Later on, we will show more, namely that $\EE T^\beta<\infty$
661: for all $0<\beta<\alpha$; see Proposition \ref{prop:fieldmoment}).
662:
663: We will also have that $T_k\to T$ a.s.\ as $k\to\infty$,
664: since, for all $k$,
665: \begin{align*}
666: 0\leq T-T_k &=\sup_{A\in\cC} \sum_{i\in A} M_i - \sup_{A\in\cC} \sum_{i\in A,
667: i\leq k} M_i \\
668: &\leq \sup_{A\in\cC} \sum_{i\in A, i>k} M_i \\
669: &= S_k.
670: \end{align*}
671:
672: The convergence in Theorem \ref{Ttheorem} will then follow from the
673: following two results, which provide control over $T_k - \tT^{(n)}_k$
674: and $\tT^{(n)}_k - \tT^{(n)}$ for appropriate $k$:
675:
676: \begin{proposition}
677: \label{coupledprop}
678: Let $\epsilon>0$ and $k$ be fixed. Then for all
679: $n$ sufficiently large,
680: say $n\geq N_k(\epsilon)$, there is a coupling
681: of the continuous model and the discrete model
682: indexed by $n$ under which
683: \begin{gather}
684: \label{coupled1}
685: \PP\left(\sum_{i=1}^k \left| M_i - \tM^{(n)}_i \right| > \epsilon
686: \right) \leq \epsilon, \\
687: \label{coupled2}
688: \PP\left( \sum_{i=1}^k \left\| Y_i - Y^{(n)}_i \right\| > \epsilon
689: \right) \leq \epsilon, \\
690: \label{coupled3}
691: \PP\left( \cC_k^{(n)}\neq \cC_k \right) \leq \epsilon.
692: \end{gather}
693: \end{proposition}
694:
695: \begin{proposition}
696: \label{Aprop}
697: Let $\epsilon>0$. Then for $k$ sufficiently large,
698: \[
699: \PP\left(\tS_k^{(n)}>\epsilon\right)\leq \epsilon
700: \]
701: for all $n$.
702: \end{proposition}
703:
704:
705: \noindent\textbf{Proof of Lemma \ref{Slemma}}:
706:
707: First, we define $L_i=\sup_{A\in\cC}\left|A\cap\{1,\dots,i\}\right|$.
708: $L_i$ is the largest number of the points $Y_1,\dots,Y_i$ (the
709: locations of the $i$ largest weights) that can be included in an
710: increasing path. Note that the collection $(L_i)$ is independent of
711: the collection $(M_i)$.
712: $L_i$ has the distribution of the ``longest increasing subsequence''
713: of a random permutation of length $i$. In particular, there is a
714: constant $c$ such that, for all $i$, $\EE L_i \leq c\sqrt{i}$
715: and $\EE L_i^2 \leq c i$; also,
716: $L_i/\sqrt{i}\to 2$ in distribution.
717: See for example \cite{AldDiapatience} for a survey.
718:
719: We will also write $U_k=\sum_{i=k+1}^\infty L_i(M_i-M_{i+1})$ for each
720: $k\geq 0$. Fix $A\in\cC$, and define $R_i=|A\cap\{1,\dots,i\}|$.
721: Then $I(i\in A)=R_i-R_{i-1}$, and by definition $R_i\leq L_i$.
722:
723: We have
724: \begin{align*}
725: \sum_{i\in A, i>k} M_i
726: &=\lim_{n\to\infty}\sum_{i\in A, k<i\leq n} M_i
727: \\
728: &=\lim_{n\to\infty}\sum_{i=k+1}^n M_i I(i\in A)
729: \\
730: &=\lim_{n\to\infty}\sum_{i=k+1}^n M_i(R_i-R_{i-1})
731: \\
732: &=\lim_{n\to\infty}
733: \left[
734: -M_{k+1}R_k + \sum_{i=k+1}^{n-1} R_i(M_i-M_{i+1})
735: +M_n R_n
736: \right]
737: \\
738: &\leq
739: \lim_{n\to\infty}
740: \sum_{i=k+1}^{n-1} R_i(M_i-M_{i+1})
741: +\liminf_{n\to\infty} M_n R_n
742: \\
743: &\leq
744: \lim_{n\to\infty}
745: \sum_{i=k+1}^{n-1} L_i(M_i-M_{i+1})
746: +\liminf_{n\to\infty} M_n L_n
747: \\
748: &=
749: U_k +\liminf_{n\to\infty} M_n L_n.
750: \end{align*}
751:
752: % Changed liminf to limsup above. And back again.
753:
754: Now
755: $\liminf_{n\to\infty} M_n L_n = 0$ a.s.;
756: this follows, for example, since
757: (by the law of large numbers) $M_n \sim n^{-1/\alpha}$ a.s.\
758: (with $\alpha<2$),
759: and since $L_n/\sqrt{n}$ converges in distribution to a constant.
760:
761: Since the inequality above holds for any $A\in\cC$,
762: we therefore have that $S_k\leq U_k$ a.s., for any $k$.
763: To conclude the proof, we will show that with probability 1,
764: $U_k<\infty$ for all $k$, and $U_k\to 0$ as $k\to\infty$.
765: (In fact, as soon as $U_1<\infty$, we necessarily
766: have that $U_k\to0$ as $k\to\infty$,
767: since the quantity $U_k$ is the ``remainder'' from
768: index $k+1$ onwards in the infinite sum $U_1$;
769: if the infinite sum $U_1$ converges, then by definition
770: these remainders tend to 0).
771:
772: Hence it's enough that $U_k<\infty$ a.s., for all $k$.
773: Specifically, we'll show that
774: $\EE U_k$ is finite whenever $k>1/\alpha$.
775: Then certainly $U_k<\infty$ a.s.\ for such $k$, and
776: in fact $U_r<\infty$ a.s.\ for all $r$, since if $r<k$,
777: $U_r-U_k$ is the sum of only finitely many terms.
778:
779: By independence of the collections $(L_i)$ and $(M_i)$,
780: \begin{align}
781: \nonumber
782: \EE U_k
783: &=\sum_{i=k+1}^\infty \EE L_i(\EE M_i - \EE M_{i+1})
784: \\
785: \label{Sbound}
786: &\leq
787: \sum_{i=k+1}^\infty ci^{1/2}(\EE M_i - \EE M_{i+1}).
788: \end{align}
789:
790: Now $M_r$ has the distribution of $(V_r)^{-1/\alpha}$,
791: where $V_r$ has Gamma($r,1$) distribution. We then obtain
792: \begin{align}
793: \EE M_r
794: \nonumber
795: &=\int_0^\infty \frac{1}{\Gamma(r)} v^{r-1}e^{-v}v^{-1/\alpha} dv
796: \\
797: \label{Gammacalc}
798: &=\Gamma\big(r-1/\alpha\big)/\Gamma(r).
799: \end{align}
800: Using the identity $\Gamma(z+1)=z\Gamma(z)$ and the
801: fact that the gamma function is log convex,
802: one has
803: \begin{equation}
804: \label{Gammafact}
805: (x-1)^a\leq \frac{\Gamma(x+a)}{\Gamma(x)} \leq (x+a)^a
806: \end{equation}
807: for $x>1$, $a<0$.
808: Then
809: \begin{align*}
810: \EE M_r-\EE M_{r+1}
811: &=\frac{\Gamma(r-1/\alpha)}{\Gamma(r)}-\frac{\Gamma(r+1-1/\alpha)}{\Gamma(r+1)}
812: \\
813: &=\frac{\Gamma(r-1/\alpha)\left[1-\frac{r-1/\alpha}{r}\right]}{\Gamma(r)}
814: \\
815: &=\frac{1}{\alpha r}\frac{\Gamma(r-1/\alpha)}{\Gamma(r)}
816: \\
817: &\leq \frac{1}{\alpha r}\big(r-1/\alpha-1)^{-1/\alpha}.
818: \end{align*}
819:
820: Returning to (\ref{Sbound}), we have
821: \[
822: \EE U_k\leq \frac{c}{\alpha}\sum_{i=k+1}^\infty i^{-1/2}
823: \big(i-1/\alpha-1)^{-1/\alpha},
824: \]
825: which is finite for all $k>1/\alpha$ (since $\alpha<2$).$\hfill\Box$
826:
827: \medskip
828:
829: \noindent\textbf{Proof of Theorem \ref{Ttheorem}}:
830: We will find a coupling of $\tilde{T}^{(n)}$ and $T$ for each $n$
831: such that $\tilde{T}^{(n)}-T\to 0$ in probability as $n\to\infty$.
832:
833: For each $n\in\NN$, define $k_n=\max\{k:n\geq N_k(1/k)\}$.
834:
835: Then $k_n\to\infty$ as $n\to\infty$, and, for all $n$,
836: $n\geq N_{k_n}(1/k_n)$.
837: Hence from Propositions \ref{coupledprop} and \ref{Aprop}
838: and from Lemma \ref{Slemma},
839: there are couplings such that, as $n\to\infty$,
840: \begin{gather}
841: \label{conv}
842: \sum_{i=1}^{k_n} \left| M_i - \tM^{(n)}_i \right| \to 0
843: \\
844: \nonumber
845: \sum_{i=1}^{k_n} \left\| Y_i - Y^{(n)}_i \right\| \to 0
846: \\
847: \nonumber
848: \tS^{(n)}_{k_n}\to 0
849: \\
850: \nonumber
851: S_{k_n}\to 0
852: \\
853: \intertext{in probability, and}
854: \nonumber
855: \PP\left( \cC_{k_n}^{(n)}\neq \cC_{k_n} \right) \to 0.
856: \end{gather}
857:
858: Now
859: \[ T-\tT^{(n)} = \left( T-T_{k_n} \right) + \left( T_{k_n} -
860: \tT^{(n)}_{k_n} \right) + \left( \tT^{(n)}_{k_n} - \tT^{(n)}. \right)
861: \]
862: We have $|T-T_{k_n}|\leq S_{k_n}$ and $|\tT^{(n)}_{k_n}-\tT^{(n)}|\leq
863: \tS^{(n)}_{k_n}$, so to show that the LHS converges to 0 in
864: probability as desired, it remains to show that $\left(
865: T_{k_n}-\tT^{(n)}_{k_n} \right)\to 0$ in probability.
866:
867: We have
868: \[
869: T_{k_n}=\max_{A\in\cC_{k_n}} \sum_{i\in A} M_i
870: \qquad
871: \text{ and }
872: \qquad
873: \tT^{(n)}_{k_n}=\max_{A\in\cC^{(n)}_{k_n}} \sum_{i\in A} \tM^{(n)}_i,
874: \]
875: so if $\cC_{k_n}=\cC^{(n)}_{k_n}$, then
876: \[
877: \left| T_{k_n}-\tT^{(n)}_{k_n} \right|
878: \leq
879: \sum_{i=1}^{k_n}\left| M_i-\tM^{(n)}_i \right|.
880: \]
881: Since $\PP\left(\cC_{k_n}\neq\cC^{(n)}_{k_n}\right)\to 0$
882: and $\sum_{i=1}^{k_n}\left| M_i-\tM^{(n)}_i \right|\to 0$
883: in probability, we are done.\enpf
884:
885: To complete the proof it remains to prove
886: Propositions \ref{coupledprop} and \ref{Aprop}.
887:
888: \subsection{Convergence of $\tT^{(n)}_k$ to $T_k$}
889:
890: \noindent\textbf{Proof of Proposition \ref{coupledprop}}:
891:
892: We have
893: $(\tM^{(n)}_1,\dots,\tM^{(n)}_k,Y^{(n)}_1,\dots,Y^{(n)}_k)
894: \to
895: (M_1,\dots,M_k,Y_1,\dots,Y_k)$ in distribution as $n\to\infty$.
896: By the Skorohod Representation Theorem, we can define
897: all the variables on the same space in such a way that
898: the convergence occurs almost surely.
899: Then indeed (\ref{coupled1}) and (\ref{coupled2}) must hold for
900: large enough $n$.
901:
902: Note that since the variables $Y_i$ are i.i.d.\ uniform
903: on $[0,1]^2$, there are almost surely no two $i$ and $j$
904: such that $Y_i(d)=Y_j(d)$ for $d=1$ or $2$.
905:
906: Thus if we perturb the point $(Y_1,\dots, Y_k)$ by a small
907: enough amount, the orderings of all the coordinates
908: remain the same, and the set $\cC_k$ of increasing paths is unchanged.
909: In fact, if
910: \[
911: \max_{1\leq i\leq k} \|Y_i-Y^{(n)}_i\|
912: \leq
913: \frac{1}{2}\min_{1\leq i,j \leq k, i\neq j} \min_{d=1,2} |Y_i(d)-Y_j(d)|,
914: \]
915: then $\cC^{(n)}_k=\cC_k$. Since we have $Y^{(n)}_i\to Y_i$
916: a.s.\ on the joint probability space for all $1\leq i\leq k$,
917: we then have $\cC^{(n)}_k=\cC_k$ eventually,
918: with probability 1.
919: Thus (\ref{coupled3}) must also hold for all large enough $n$,
920: as desired. $\hfill\Box$
921:
922:
923: \subsection{Convergence of $\tT^{(n)}-\tT^{(n)}_k$ to $0$}
924:
925: Our aim in this section is to prove Proposition \ref{Aprop}.
926:
927: Define the ``good event'' $\cB^{(n)}_k$:
928: \begin{equation}
929: \label{Bdef}
930: \cB^{(n)}_k =\left\{
931: F^{-1}\left(1-\frac{2r}{n^2}\right)
932: \leq
933: M^{(n)}_r
934: \leq
935: F^{-1}\left(1-\frac{1}{n^2}\right)
936: \text{ for all } k<r\leq n^2
937: \right\}.
938: \end{equation}
939:
940: \begin{lemma}
941: \label{Blemma}
942: $\PP\left(\cB^{(n)}_k\right)\to 1$ as $k\to\infty$, uniformly in $n$.
943: \end{lemma}
944:
945: \noindent\textit{Proof:}
946: \begin{align*}
947: \PP\left(\cB^{(n)}_k \text{ fails }\right)
948: &\leq \PP\left( M^{(n)}_{k+1}> F^{-1}\left(1-\frac{1}{n^2}\right)
949: \right) +\sum_{r=k+1}^{\lfloor n^2/2 \rfloor}
950: \PP\left( M^{(n)}_r< F^{-1}\left(1-\frac{2r}{n^2}\right) \right)
951: \\
952: &= \PP\left(\text{Binomial}(n^2,1/n^2)\geq k+1\right)
953: +\sum_{r=k+1}^{\lfloor n^2/2 \rfloor}
954: \PP\left(\text{Binomial}(n^2, 2r/n^2) < r\right)
955: \\
956: &\leq\frac{1}{k+1} + \sum_{r=k+1}^{\lfloor n^2/2 \rfloor}
957: 2\exp\left(-\frac{r}{12}\right),
958: \end{align*}
959: using Markov's inequality for the first term and
960: an estimate from Corollary 2.3 of \cite{JLRbook} for the second.
961: The RHS tends to 0 as $k\to\infty$, uniformly in $n$, as
962: required.$\hfill\Box$
963:
964: \medskip
965:
966: Now, we will prove a bound on the expectation
967: of $\tS^{(n)}_k$
968: in terms of the ``order statistics''
969: $M^{(n)}_r$.
970: Define
971: \[
972: L^{(n)}_i= \max_{A\in\cC^{(n)}} \left| A\cap\{1,2,\dots,i\} \right|.
973: \]
974: Recall that $Y^{(n)}_r\in\{1,2,\dots,n\}^2$ is
975: the location of the $r$th largest weight, $M^{(n)}_r$.
976: Thus, $L^{(n)}_i$ is the maximum number of the
977: points $Y^{(n)}_1,\dots,Y^{(n)}_i$ that can be included
978: in an increasing path.
979:
980: %Lemma \ref{pathsizelemma} tells us that
981: %$\EE L^{(n)}_m\leq c\sqrt{m}$ for all $m\leq n^2$
982: %(since the distribution of $\{Y^{(n)}_1,\dots,Y^{(n)}_m\}$
983: %is uniform over the subsets of $\{1,\dots,n\}$ of size $m$).
984:
985: Note that the collection $(L^{(n)}_r)_{1\leq r\leq n^2}$
986: is a function of the values $Y^{(n)}_r$ alone;
987: in particular it is independent of the weights $M^{(n)}_r$
988: and of the events $B^{(n)}_r$.
989:
990: \begin{lemma}
991: \label{Lboundlemma}
992: There is a constant $c$ independent of $m$ and $n$ such
993: that $\EE L^{(n)}_m \leq c\sqrt{m}$, whenever $1\leq m\leq n^2$.
994: \end{lemma}
995:
996: \noindent\textit{Proof:}
997: The distribution of
998: $\{Y^{(n)}_1,\dots,Y^{(n)}_m\}$
999: is uniform over the subsets of $\{1,\dots,n\}^2$ of size $m$,
1000: and
1001: $L^{(n)}_r$ is the maximum number of the
1002: points $Y^{(n)}_1,\dots,Y^{(n)}_r$ that can be included
1003: in an increasing path.
1004:
1005: We compare this with the last-passage percolation
1006: problem in $\{1,\dots,n\}^2$ with i.i.d.\
1007: Bernoulli($p$) weights.
1008:
1009: We have the following representation for the expectation
1010: of the passage time for such a problem:
1011: \begin{equation}
1012: \label{dec}
1013: \EE_{\Ber(p)} T(n,n)=\sum_{r} \PP(B_{p,n^2}=r)\EE L^{(n)}_r,
1014: \end{equation}
1015: where
1016: $B_{p,n^2}$ has Binomial($p,n^2$) distribution,
1017: since, conditional on the event
1018: that exactly $r$ of the $n^2$ weights
1019: have value 1, the set of positions of the weights
1020: with value 1 is uniformly distributed among all the
1021: subsets of $\{1,\dots,n\}^2$ of size $r$.
1022:
1023: Proposition 2.2 of \cite{animals} shows that there
1024: exists a constant $c_1$ such that
1025: \begin{equation}
1026: \label{animalsbound}
1027: \EE_{\Ber(p)} T(n,n)\leq c_1 p^{1/2}n
1028: \end{equation}
1029: for all $n$, $p$.
1030:
1031: Given $1\leq m\leq n^2$, set $p=\min(2m/n^2,1)$.
1032: Using (\ref{dec}), (\ref{animalsbound}) and the
1033: fact that $\EE L^{(n)}_r$ is increasing in $r$,
1034: we obtain
1035: \begin{align*}
1036: \PP(B_{p,n^2}\geq m)\EE L^{(n)}_m
1037: &\leq \sum_{r} \PP(B_{p,n^2}=r)\EE L^{(n)}_r
1038: \\
1039: &\leq c_1 \left(\min(2m/n^2,1)\right)^{1/2} n
1040: \\
1041: &\leq c_2\sqrt{m}.
1042: \end{align*}
1043:
1044: To complete the proof, it then suffices to bound
1045: $\PP(B_{p,n^2}\geq m)$ away from 0 uniformly in $1\leq m\leq n^2$.
1046:
1047: If $m\geq n^2/2$ then $p=1$ and $\PP(B_{p,n^2}\geq m)=1$.
1048:
1049: For $1\leq m\leq n^2/2$, we have $pn^2/2=m$,
1050: and we use the estimate
1051: \[
1052: \PP(B_{p,n^2}<pn^2/2)\leq \exp(-pn^2/8),
1053: \]
1054: (see for example Theorem 2.1 of \cite{JLRbook}),
1055: to give
1056: $\PP(B_{p,n^2}\geq m)\geq 1-\exp(-1/8)$ uniformly
1057: in $1\leq m\leq n^2/2$
1058: as desired.$\hfill\Box$
1059:
1060: \begin{lemma}
1061: \label{differencelemma}
1062: \[
1063: \EE\left(\tS^{(n)}_k;\cB^{(n)}_k\right)
1064: \leq
1065: c(k+1)^{1/2}\EE \left(\tM^{(n)}_{k+1};\cB^{(n)}_k\right)
1066: + c\sum_{r=k+2}^{n^2} r^{-1/2} \EE \left(\tM^{(n)}_r;\cB^{(n)}_k\right).
1067: \]
1068: \end{lemma}
1069:
1070: \noindent\textit{Proof:}
1071: The argument is similar to the proof of Lemma \ref{Slemma}.
1072:
1073: Let $\tA$ achieve the max in the definition of $\tS^{(n)}_k$,
1074: so that $\tS^{(n)}_k=\sum_{i\in\tA, i>k}\tM^{(n)}_i$.
1075:
1076: Define
1077: $R_i=|\tA\cap\{1,2,\dots,i\}|$ for each $i$.
1078: Then $R_i-R_{i-1}=I(i\in\tA)$,
1079: and by definition $R_i\leq L_i^{(n)}$ for each $i$.
1080: We then have
1081: \begin{align*}
1082: \tS^{(n)}_k
1083: &=
1084: \sum_{i\in\tA, i>k} \tM^{(n)}_i
1085: \\
1086: &= \sum_{i=k+1}^{n^2} \tM^{(n)}_i\left(R_i-R_{i-1}\right)
1087: \\
1088: &= -R_k \tM^{(n)}_{k+1}
1089: +\sum_{i=k+1}^{n^2-1} R_i\left(\tM^{(n)}_i-\tM^{(n)}_{i+1}\right)
1090: +R_{n^2} \tM^{(n)}_{n^2}
1091: \\
1092: &\leq
1093: \sum_{i=k+1}^{n^2-1} L^{(n)}_i\left(\tM^{(n)}_i-\tM^{(n)}_{i+1}\right)
1094: +L^{(n)}_{n^2} \tM^{(n)}_{n^2},
1095: \end{align*}
1096: since $R_i^{(n)}\leq L_i^{(n)}$ and $\tM_i^{(n)}\geq \tM_{i+1}^{(n)}$.
1097:
1098: We now take expectations, restricted to the event
1099: $\cB^{(n)}_k$, using
1100: Lemma \ref{Lboundlemma} and
1101: the independence of the $L^{(n)}_r$
1102: from the $\tM^{(n)}_r$:
1103: %\begin{multline*}
1104: \begin{align*}
1105: \EE\left(\tS^{(n)}_k;\cB^{(n)}_k\right)
1106: %\\
1107: %\begin{aligned}
1108: &\leq
1109: \sum_{i=k+1}^{n^2-1}
1110: \EE L^{(n)}_i
1111: \left[\EE\left(\tM^{(n)}_i;\cB^{(n)}_k\right)
1112: -\EE\left(\tM^{(n)}_{i+1};\cB^{(n)}_k\right)\right]
1113: +\EE L^{(n)}_{n^2}
1114: \EE\left(\tM^{(n)}_{n^2};\cB^{(n)}_k\right)
1115: \\
1116: &\leq
1117: \sum_{i=k+1}^{n^2-1}
1118: c\sqrt{i}
1119: \left[\EE\left(\tM^{(n)}_i;\cB^{(n)}_k\right)
1120: -\EE\left(\tM^{(n)}_{i+1};\cB^{(n)}_k\right)\right]
1121: +
1122: cn
1123: \EE L^{(n)}_{n^2}
1124: \EE\left(\tM^{(n)}_{n^2};\cB^{(n)}_k\right)
1125: \\
1126: &=
1127: c\sqrt{k+1}\EE \left(\tM^{(n)}_{k+1};\cB^{(n)}_k\right)
1128: +
1129: c\sum_{i=k+2}^{n^2}
1130: \left(\sqrt{i}-\sqrt{i-1}\right)\EE \left(\tM^{(n)}_i;\cB^{(n)}_k\right)
1131: \\
1132: &\leq
1133: c\sqrt{k+1}\EE \left(\tM^{(n)}_{k+1};\cB^{(n)}_k\right)
1134: +
1135: c\sum_{i=k+2}^{n^2}
1136: i^{-1/2}
1137: \EE \left(\tM^{(n)}_i;\cB^{(n)}_k\right),
1138: \end{align*}
1139: %\end{aligned}
1140: %\end{multline*}
1141: since $i^{1/2}-(i-1)^{1/2}\leq i^{-1/2}$.
1142: This is the required result.$\hfill\Box$
1143:
1144: \medskip
1145:
1146: The next lemma gives an estimate
1147: on the tail behaviour of the weight distribution
1148: using the regular variation condition.
1149: Note that if the weight distribution were Pareto($\alpha$),
1150: then $F^{-1}(u)=(1-u)^{-1/\alpha}$,
1151: and one then has exactly
1152: $F^{-1}(u_1)=
1153: F^{-1}(u_0)\left[
1154: (1-u_1)/(1-u_0)
1155: \right]^{-1/\alpha}$ for any $u_0$, $u_1$.
1156:
1157: \begin{lemma}
1158: \label{regboundlemma}
1159: For any $\delta>0$, there exists $U(\delta)<1$ such that
1160: for all $u_0, u_1$ with $U(\delta) \leq u_1 \leq u_0$,
1161: \[
1162: F^{-1}(u_1)\leq
1163: 2F^{-1}(u_0)\left(
1164: \frac{1-u_1}{1-u_0}
1165: \right)^{-\frac{1}{\alpha}+\delta}.
1166: \]
1167: \end{lemma}
1168:
1169: \noindent\textit{Proof:}
1170: Fix $s>1$ sufficiently small that $s^{1/\alpha-\delta}<2$.
1171:
1172: From the fact that the tail of $F$ is regularly
1173: varying with index $\alpha$, the following property holds:
1174: if $u_1$ is sufficiently close to $1$ (at least $U(\delta)$, say),
1175: then for all $(1-u_0)<(1-u_1)/s$,
1176: \[
1177: \frac{F^{-1}(u_1)}{F^{-1}(u_0)} < s^{-\frac{1}{\alpha}+\delta}.
1178: \]
1179:
1180: Iterating, one obtains that if $U(\delta)\leq u_1\leq u_0$, then
1181: \begin{align*}
1182: \frac{F^{-1}(u_1)}{F^{-1}(u_0)}
1183: &<
1184: \left(
1185: s^{-\frac{1}{\alpha}+\delta}
1186: \right)^{\left\lfloor \log_s \left((1-u_1)/(1-u_0)\right)\right\rfloor}
1187: \\
1188: &=
1189: \exp\left[
1190: \left(-\frac{1}{\alpha}+\delta\right)(\log s)
1191: \genfrac\lfloor\rfloor{}{}{\log\left((1-u_1)/(1-u_0)\right)}{\log s}
1192: \right]
1193: \\
1194: &\leq
1195: \exp\left[
1196: \left(-\frac{1}{\alpha}+\delta\right)
1197: \left(\log\frac{1-u_1}{1-u_0}-\log s\right)
1198: \right]
1199: \\
1200: &=
1201: \genfrac(){}{}{1-u_1}{1-u_0}^{-\frac{1}{\alpha}+\delta}
1202: s^{\frac{1}{\alpha}-\delta}
1203: \\
1204: &\leq
1205: 2\genfrac(){}{}{1-u_1}{1-u_0}^{-\frac{1}{\alpha}+\delta}
1206: \end{align*}
1207: as required.$\hfill\Box$
1208:
1209: \medskip
1210:
1211: Finally we use this tail estimate to control
1212: the expectation of the variables $\tM^{(n)}_r$,
1213: restricted to the ``good set'' $\cB^{(n)}_k$.
1214:
1215: \begin{lemma}
1216: \label{EMlemma}
1217: Let $\delta>0$. Then there exist $c_0$, $c_1$ and $c_2>0$ such that
1218: \[
1219: \EE\left(
1220: \tM^{(n)}_r; \cB^{(n)}_k
1221: \right)
1222: \leq
1223: c_0 r^{-\frac{1}{\alpha}+\delta}
1224: + c_1 a_{n^2}^{-1} I(r\geq c_2 n^2)
1225: \]
1226: for all $n$, $k$, $r$ satisfying
1227: $2(1+1/\alpha)< k < r \leq n^2$.
1228: \end{lemma}
1229:
1230: \noindent\textit{Proof:}
1231: Let $U(\delta)$ be as in Lemma \ref{regboundlemma},
1232: and set $c_2 = (1-U(\delta))/2$.
1233: %%$c_2=\sqrt{2/(1-U(\delta))}$.
1234: Then
1235: $r<c_2 n^2 \Leftrightarrow 1-2r/n^2 > U(\delta)$.
1236:
1237: Note that if
1238: \[
1239: \max\left\{U(\delta),1-\frac{2r}{n^2}\right\} \leq u \leq 1-\frac{1}{n^2},
1240: \]
1241: then, by Lemma \ref{regboundlemma},
1242: \begin{align}
1243: \nonumber
1244: F^{-1}(u)
1245: &\leq
1246: 2F^{-1}\left(1-\frac{1}{n^2}\right)
1247: \left[n^2(1-u)\right]^{-\frac{1}{\alpha}+\delta}
1248: \\
1249: \nonumber
1250: &= 2 a_{n^2} n^{-2/\alpha}(1-u)^{-1/\alpha}\left[n^2(1-u)\right]^\delta
1251: \\
1252: \label{powerbound}
1253: &\leq
1254: 2 a_{n^2} n^{-2/\alpha}(1-u)^{-1/\alpha}(2r)^\delta.
1255: \end{align}
1256:
1257: Since $r>k$, we have that
1258: \[
1259: \cB^{(n)}_k\subseteq
1260: \left\{
1261: F^{-1}\left(1-\frac{2r}{n^2}\right)
1262: \leq
1263: M^{(n)}_r
1264: \leq
1265: F^{-1}\left(1-\frac{1}{n^2}\right)
1266: \right\}.
1267: \]
1268:
1269:
1270: Hence
1271: \begin{align}
1272: \nonumber
1273: \EE\Big(\tM^{(n)}_r ; &\cB^{(n)}_k\Big)
1274: \\
1275: &\leq a_{n^2}^{-1} \EE\left(M^{(n)}_r ;
1276: F^{-1}\left(1-\frac{2r}{n^2}\right)
1277: \leq M^{(n)}_r \leq F^{-1}\left(1-\frac{1}{n^2}\right)
1278: \right)
1279: \\
1280: \nonumber
1281: &= a_{n^2}^{-1} \EE\left(F^{-1}\left(U^{(n)}_r\right) ;
1282: 1-\frac{2r}{n^2} \leq U^{(n)}_r \leq 1-\frac{1}{n^2} \right)
1283: \\ \nonumber &= a_{n^2}^{-1}
1284: \int_{1-2r/n^2}^{1-1/n^2} F^{-1}(u) f_{r;n^2}(u) du
1285: \\ \nonumber &= a_{n^2}^{-1}
1286: I\left\{1-\frac{2r}{n^2} \leq U(\delta)\right\}F^{-1}(U(\delta))
1287: +a_{n^2}^{-1}
1288: \int_{\max\left\{U(\delta),1-2r/n^2\right\}}^{1-1/n^2}
1289: F^{-1}(u) f_{r;n^2}(u) du
1290: \\
1291: \nonumber
1292: &\leq
1293: c_1 a_{n^2}^{-1}
1294: I\left(r\geq c_2 n^2\right)
1295: +\int_{\max\left\{U(\delta),1-2r/n^2\right\}}^{1-1/n^2}
1296: 2 n^{-2/\alpha}(1-u)^{-1/\alpha}(2r)^\delta
1297: f_{r;n^2}(u) du
1298: \\
1299: \label{midstep}
1300: &\leq
1301: c_1 I\left(r\geq c_2 n^2\right)
1302: +c_3 a_{n^2} n^{-2/\alpha}r^{\delta}
1303: \int_0^1 (1-u)^{-1/\alpha}f_{r;n^2}(u) du,
1304: \end{align}
1305: where $c_1=F^{-1}(U(\delta))$ and $c_3=2^{1+\delta}$,
1306: and where $f_{r;n^2}$ is the density function
1307: of the $r$th largest from an i.i.d.\ sample of size $n^2$ from
1308: the uniform distribution on $[0,1]$.
1309:
1310: Now
1311: \[
1312: f_{r;n^2}(u)=
1313: \frac{\Gamma(n^2+1)}{\Gamma(n^2-r+1)\Gamma(r)}
1314: (1-u)^{r-1}u^{n^2-r},
1315: \]
1316: and, since $r-1-1/\alpha>0$, one then has
1317: \begin{align}
1318: \nonumber
1319: \int_0^1 (1-u)^{-1/\alpha}f_{r;n^2}(u) du
1320: &=
1321: \frac{\Gamma(n^2+1)}{\Gamma(n^2-r+1)\Gamma(r)}
1322: \int_0^1 (1-u)^{r-1-1/\alpha}u^{n^2-r} du
1323: \\
1324: \nonumber
1325: &=
1326: \frac{\Gamma(n^2+1)}{\Gamma(n^2-r+1)\Gamma(r)}
1327: \genfrac(){}{}
1328: {\Gamma(n^2+1-1/\alpha)}
1329: {\Gamma(n^2-r+1)\Gamma(r-1/\alpha)}^{-1}
1330: \\
1331: \label{Gammaid}
1332: &=
1333: \frac{\Gamma(n^2+1)}{\Gamma(n^2+1-1/\alpha)}
1334: \frac{\Gamma(r-1/\alpha)}{\Gamma(r)}.
1335: \end{align}
1336:
1337: Using (\ref{Gammafact}),
1338: the RHS of (\ref{Gammaid}) is bounded above by
1339: \[
1340: \frac{(n^2+1)^{1/\alpha}}{(r-1-1/\alpha)^{1/\alpha}}.
1341: \]
1342: Since $r>2(1+1/\alpha)$, this is in turn
1343: no greater than $\left(4n^2/r\right)^{1/\alpha}$.
1344: Inserting this into (\ref{midstep}) gives the desired
1345: result.$\hfill\Box$
1346:
1347: \medskip
1348:
1349: \noindent\textbf{Proof of Proposition \ref{Aprop}:}
1350: We may assume that $k\leq n^2$,
1351: since if $k>n^2$ then $\tS^{(n)}_k=0$.
1352:
1353: Fix $\epsilon>0$, and fix some $\delta<\frac{1}{\alpha}-\frac{1}{2}$.
1354: Using Markov's inequality, we have
1355: \[
1356: \PP\left(\tS^{(n)}_k>\epsilon\right)
1357: \leq
1358: \PP(\cB^{(n)}_k \text{ fails })
1359: + \epsilon^{-1}
1360: \EE
1361: \left(\tS^{(n)}_k; \cB^{(n)}_k \right).
1362: \]
1363:
1364: By Lemma \ref{Blemma}, the first term tends to 0 as $k\to\infty$,
1365: uniformly in $n$. For the second term, Lemmas \ref{differencelemma}
1366: and \ref{EMlemma} combine to give
1367: \begin{multline*}
1368: \EE
1369: \left(\tS^{(n)}_k; \cB^{(n)}_k \right)
1370: \leq
1371: cc_0(k+1)^{-\frac{1}{\alpha}+\frac{1}{2}+\delta}
1372: +
1373: cc_0\sum_{r=k+2}^{n^2} r^{-\frac{1}{\alpha}-\frac{1}{2}+\delta}
1374: \\
1375: +
1376: cc_1(k+1)^{\frac{1}{2}}a_{n^2}^{-1}
1377: +
1378: cc_1 a_{n^2}^{-1} \sum_{c_2 n^2 \leq r \leq n^2} r^{-\frac{1}{2}}.
1379: \end{multline*}
1380: From the choice of $\delta$ and the fact that $a_{n^2}/n\to \infty$ as
1381: $n\to\infty$,
1382: one obtains that
1383: all four terms on the RHS tend to 0 as $k\to\infty$ uniformly in $n$
1384: such that $k\leq n^2$.
1385:
1386: Hence indeed
1387: $\PP\left(\tS^{(n)}_k>\epsilon\right)<\epsilon$
1388: for all large enough $k$, uniformly in $n$
1389: such that $k\leq n^2$,
1390: as required.$\hfill\Box$
1391:
1392: \section{Path convergence}
1393: \label{pathsection}
1394:
1395: In this section we will state and prove results describing
1396: the convergence of the distribution of the
1397: optimal paths for the discrete
1398: models to that for the limiting continuous model.
1399:
1400: First we note that the optimal path for the continuous model
1401: is well-defined:
1402:
1403: \begin{proposition}
1404: \label{uniquenessprop}
1405: With probability 1,
1406: there exists a unique $A^*\in\cC$ such that
1407: $T=\sum_{i\in A^*} M_i$.
1408: \end{proposition}
1409:
1410: We will also define $A^{(n)*}$
1411: as the set that achieves the maximum in
1412: \[
1413: T^{(n)}=\max_{A\in\cC^{(n)}}\sum_{i\in A} M^{(n)}_i.
1414: \]
1415: (or in the equivalent expression for $\tT^{(n)}$).
1416: (Since the weight distribution is assumed to be continuous,
1417: this optimal set is almost surely unique).
1418:
1419: It will be useful to extend the sequences
1420: $(Y^{(n)}_i)_i$ and $(\tM^{(n)}_i)_i$ to all $i\in\NN$
1421: (rather than only $i\leq n^2$);
1422: for example, we can put $\tM^{(n)}_i=0$, $Y^{(n)}_i=(0,0)$
1423: for all $i\geq n^2$. We will consider always the
1424: product topology when looking at convergence
1425: of such infinite sequences.
1426:
1427: We also use the product topology on $\cS$, the set of subsets of $\NN$.
1428: Thus, given $A$ and a sequence $A_k$ in $\cS$, we have $A_k\to A$ if, for
1429: every $m$, $A_k \cap \{1,\dots,m\}$ is equal to $A \cap \{1,\dots,m\}$
1430: for all large enough $k$. One has easily that any sequence $A_k$ has
1431: at least one limit point, and also that if $A_k\in\cC$ for each $k$
1432: then every limit point is also in $\cC$ (since if the limit point
1433: contains $i$ and $j$, then $i$ and $j$ are in $A_k$ for some $k$, and
1434: hence $Y_i\sim Y_j$).
1435:
1436: The following theorem is our first path convergence result.
1437: Later (in Theorem \ref{paththeorem2}) we will use
1438: it to prove a more direct convergence result
1439: concerning the optimal paths viewed as random subsets of $[0,1]^2$.
1440:
1441: \begin{theorem}
1442: \label{paththeorem}
1443: $\Big(\big(Y^{(n)}_i\big)_{i\in\NN}, A^{(n)*}\Big)
1444: \to
1445: \Big(\big(Y_i\big)_{i\in\NN}, A^{*}\Big)$
1446: in distribution as $n\to\infty$.
1447: \end{theorem}
1448:
1449: Before proving Proposition \ref{uniquenessprop} and
1450: Theorem \ref{paththeorem}, we need the following fact:
1451: \begin{lemma}
1452: \label{pathlemma}
1453: With probability 1, the following holds:
1454: if $A_j$ is a sequence in $\cC$ converging to
1455: a limit $A$, then
1456: $\lim_{j\to\infty}\sum_{i\in A_j} M_i=\sum_{i\in A} M_i$.
1457: \end{lemma}
1458:
1459: \noindent
1460: [N.B.\ this result is not true in general for sequences $A_j$ in $\cS$
1461: (unless $\alpha<1$ so that $\sum M_i<\infty$ a.s.)]
1462: \medskip
1463:
1464: \noindent\textit{Proof:}
1465: If $A_j \cap \{1,\dots,m\}=A \cap \{1,\dots,m\}$ then
1466: \[ \left|\sum_{i\in A_j} M_i - \sum_{i\in A} M_i \right|
1467: \leq \sup_{\tA\in\cC} \sum_{i\in\tA,i>m} M_i =S_m. \]
1468: But $S_m\to 0$ a.s.\ as $m\to\infty$ (from Lemma \ref{Slemma}),
1469: and $A_j \cap \{1,\dots,m\}=A \cap \{1,\dots,m\}$
1470: eventually for all $m$, so we are done.$\hfill\Box$
1471:
1472: \medskip
1473:
1474: \noindent\textbf{Proof of Proposition \ref{uniquenessprop}}:
1475: Recall that
1476: \[
1477: T_k=\sup_{A\in\cC}\sum_{i\in A, i\leq k} M_i.
1478: \]
1479: Since the sum on the RHS depends only
1480: on the intersection of $A$ with $\{1,\dots,k\}$,
1481: we only need to consider the max over finitely
1482: many $A\subseteq\{1,\dots,k\}$.
1483: Thus there exists some $A^*_{k}$ which achieves the sup.
1484:
1485:
1486: %Note that if $A_k\to A$ and each $A_k$ is in $\cC$,
1487: %then also $A\in\cC$
1488: %(since if $i$, $j$ are in $A$,
1489: %then $i$ and $j$ are in $A_k$ for some $k$,
1490: %and hence $Y_i$ and $Y_j$ are compatible).
1491:
1492: We consider the sequence $A^*_k$.
1493: As observed above, this sequence has
1494: at least one limit point $A^*\in\cC$,
1495: and by Lemma \ref{pathlemma},
1496: $\sum_{i\in A^*} M_i=\lim_{k\to\infty} \sum_{i\in A^*_k} M_i
1497: =\lim_{k\to\infty} T_k = T$.
1498:
1499: Now we wish to show that in fact a unique $A^*$
1500: achieves the sum $T$.
1501:
1502: Suppose instead that there are two such optimising sets in $\cC$.
1503: Then there is some $k$ that is contained in one but not the other.
1504:
1505: In that case,
1506: \begin{align*}
1507: \sup_{A\in\cC, k\notin A} \sum_{i\in A} M_i
1508: &=
1509: \sup_{A\in\cC, k\in A} \sum_{i\in A} M_i
1510: \\
1511: &=M_k+\sup_{A\in\cC, k\in A} \sum_{i\in A, i\neq k} M_i,
1512: \end{align*}
1513: which gives
1514: \begin{equation}\label{event}
1515: M_k=
1516: \sup_{A\in\cC, k\notin A} \sum_{i\in A} M_i
1517: -
1518: \sup_{A\in\cC, k\in A} \sum_{i\in A, i\neq k} M_i.
1519: \end{equation}
1520: We will show that this event has probability 0 for each $k$;
1521: then, by countable additivity, we are done.
1522:
1523: The RHS of (\ref{event}) does not depend on $M_k$;
1524: in fact, it is a function of the collection
1525: $\big( (M_i)_{i\in\NN\setminus\{k\}}, (Y_i)_{i\in\NN} \big)$.
1526: We condition on the value of this collection.
1527: Then the RHS is a constant, while the random variable
1528: $M_k$ on the LHS has a continuous distribution.
1529: (Specifically, the distribution
1530: of $M_k^{-\alpha}$ conditional on this collection
1531: is uniform on the interval
1532: $\left( M_{k-1}^{-\alpha}, M_{k+1}^{-\alpha}\right)$,
1533: since the sequence $\left(M_i^{-\alpha}\right)$
1534: forms the points of a Poisson process).
1535: Hence the event (\ref{event}) has probability 0,
1536: as required.$\hfill\Box$
1537:
1538: \medskip
1539:
1540: \noindent\textbf{Proof of Theorem \ref{paththeorem}}:
1541: We will use the same couplings as in the proof of
1542: Theorem \ref{Ttheorem}.
1543: As at (\ref{conv}), we then have that
1544: $\PP\left(
1545: \cC_{k_n}^{(n)}\neq \cC_{k_n}
1546: \right)
1547: \to 0$ as $n\to\infty$,
1548: and that all of the quantities
1549: $\sum_{i=1}^{k_n} \left| M_i - \tM^{(n)}_i \right|$,
1550: $\sum_{i=1}^{k_n} \left\| Y_i - Y^{(n)}_i \right\|$,
1551: $\tS^{(n)}_{k_n}$ and
1552: $S_{k_n}$ converge to 0
1553: in probability as $n\to\infty$.
1554:
1555: To prove Theorem \ref{paththeorem}, it will then suffice
1556: to show in addition that for any $m$,
1557: \[
1558: \PP\left(A^{(n)*}\cap\{1,\dots,m\}\neq
1559: A^{*}\cap\{1,\dots,m\}\right)\to 0
1560: \]
1561: as $n\to\infty$.
1562: For this, it's in turn enough to show that for all $r$,
1563: \begin{gather}
1564: \label{firstone}
1565: \PP\left(r\in A^*,r\notin A^{(n)*}\right)\to 0
1566: \\
1567: \label{secondone}
1568: \PP\left(r\notin A^*,r\in A^{(n)*}\right)\to 0
1569: \end{gather}
1570: as $n\to\infty$.
1571: We will show (\ref{firstone}); an analogous argument gives
1572: (\ref{secondone}).
1573:
1574: Define $T_{(-r)}=\sup_{A\in\cC, r\notin A} \sum_{i\in A} M_i$.
1575:
1576: Suppose $r\in A^*$.
1577: Then $T_{(-r)} < T$ strictly.
1578: (Otherwise, there is a
1579: sequence of members of $\cC$, none of which contain $r$,
1580: whose weight converges to $T$;
1581: then (by Lemma \ref{pathlemma})
1582: this sequence has some limit point,
1583: itself a member of $\cC$ not containing $r$,
1584: which attains the weight $T$. But this contradicts the
1585: uniqueness of $A^*$ established in Proposition
1586: \ref{uniquenessprop}).
1587:
1588: If $\cC^{(n)}_{k_n}=\cC_{k_n}$, then
1589: \begin{align}
1590: \nonumber
1591: \max_{A\in\cC^{(n)}, r\notin A} \sum_{i\in A} \tM^{(n)}_i
1592: &\leq
1593: \max_{A\in\cC^{(n)}_{k_n}, r\notin A} \sum_{i\in A} \tM^{(n)}_i
1594: +\tS^{(n)}_{k_n}
1595: \\
1596: \nonumber
1597: &\leq
1598: \max_{A\in\cC_{k_n}, r\notin A} \sum_{i\in A} \tM^{(n)}_i
1599: +\sum_{i=1}^{k_n} \left[\tM^{(n)}_i - M_i\right]_+
1600: +\tS^{(n)}_{k_n}
1601: \\
1602: \label{ineq1}
1603: &\leq
1604: T_{(-r)} +
1605: \sum_{i=1}^{k_n} \left[ \tM^{(n)}_i -M_i \right]_+
1606: + \tS^{(n)}_{k_n},
1607: \\
1608: \intertext{and similarly}
1609: \label{ineq2}
1610: \max_{A\in\cC^{(n)}, r\in A} \sum_{i\in A} \tM^{(n)}_i
1611: &\geq
1612: T -
1613: \sum_{i=1}^{k_n} \left[ M_i - \tM^{(n)}_i \right]_+
1614: - S_{k_n}.
1615: \end{align}
1616: If also $r\notin A^{(n)*}$, then
1617: \[
1618: \max_{A\in\cC^{(n)}, r\notin A} \sum_{i\in A} \tM^{(n)}_i
1619: \geq
1620: \max_{A\in\cC^{(n)}, r\in A} \sum_{i\in A} \tM^{(n)}_i,
1621: \]
1622: and using (\ref{ineq1}) and (\ref{ineq2}) we get
1623: \[
1624: T-T_{(-r)}
1625: \leq
1626: \sum_{i=1}^{k_n} \left| M_i - \tM^{(n)}_i \right|
1627: + S_{k_n} + \tS^{(n)}_{k_n}
1628: \]
1629: So altogether we obtain
1630: \begin{multline*}
1631: \PP\left(r\in A^*, r\notin A^{(n)*}\right)
1632: \\
1633: \leq
1634: \PP\left(
1635: \cC_k^{(n)}\neq \cC_k
1636: \right)
1637: +
1638: \PP\left(
1639: 0<
1640: T-T_{(-r)}
1641: \leq
1642: \sum_{i=1}^{k_n} \left| M_i - \tM^{(n)}_i \right|
1643: + S_{k_n} + \tS^{(n)}_{k_n}
1644: \right).
1645: \end{multline*}
1646: We have already observed above that the first probability on the RHS
1647: tends to 0 as $n\to\infty$. The same is true for the second probability
1648: on the RHS,
1649: since all the terms on the right of the inequality converge
1650: to 0 in probability as $n\to\infty$, while the term in the middle of
1651: the inequality does not depend on $n$.
1652: Hence $\PP\left(r\in A^*, r\notin A^{(n)*}\right)\to 0$ as $n\to\infty$,
1653: as required.\enpf
1654:
1655: We now turn to the convergence of the paths
1656: regarded as subsets of $[0,1]^2$.
1657:
1658: First let $U^*=\bigcup_{i\in A^*} Y_i \cup \{(0,0), (1,1)\}$,
1659: and take its closure $\Ubar^*$.
1660:
1661: We expect that $\Ubar^*$ is connected with
1662: probability 1, but we don't have a proof. To work around
1663: this, we will use the fact that, at least,
1664: there is a.s.\ a unique way to extend
1665: $\Ubar^*$ to a connected set while preserving the increasing
1666: path property. (Here the increasing path property of a set
1667: means that if $y$ and $y'$ are two elements of the set then $y\sim y'$).
1668:
1669: To see this, first note that if $\Ubar^*$ does ``contain jumps'',
1670: then none of these jumps can span a rectangle of non-zero area.
1671: That is, with probability 1 there is no rectangle $R$
1672: of non-zero area such that $y\sim y'$ for all $y\in\Ubar^*$
1673: and $y'\in R$.
1674:
1675: For if there were, then $R$ would certainly contain some
1676: points $Y_j$; such $j$ could be added to $A^*$,
1677: increasing the weight of the path by $M_j$; this contradicts
1678: the maximality of $A^*$.
1679:
1680: So any jumps in $\Ubar^*$ consist only of horizontal or
1681: vertical line segments. These segments can all be added to $\Ubar^*$
1682: while still preserving the increasing path property, and
1683: this gives a connected set. Conversely, any connected increasing
1684: set containing $\Ubar^*$ must ``fill in'' these jumps.
1685:
1686: Thus, define $P^*$ by setting $y\in P^*$ if:
1687: \begin{itemize}
1688: \item[(i)]$y\in\Ubar^*$, or
1689: \item[(ii)]there exists $y',y''\in\Ubar^*$ with either
1690: \begin{itemize}
1691: \item[(a)]$y'(1)=y(1)=y''(1)$, $y'(2)<y(2)<y''(2)$, or
1692: \item[(b)]$y'(2)=y(2)=y''(2)$, $y'(1)<y(1)<y''(1)$.
1693: \end{itemize}
1694: \end{itemize}
1695:
1696: This set $P^*$ (which we conjecture to be equal to
1697: the closure of $\bigcup_{i\in A^*} Y_i$ w.\ p.\ 1)
1698: provides the distributional limit we need.
1699:
1700: For each $n$, we define the object representing
1701: the optimal path in the discrete problem indexed by $n$ as follows:
1702: order the points $\{Y_i^{(n)}, i\in A^{(n)*}\}$
1703: in increasing order and join successive points by a straight line
1704: (horizontal or vertical, of length $1/n$). Call the
1705: resulting path $P^{(n)*}$.
1706:
1707: $P^*$ and $P^{(n)*}$ are regarded as subsets of $[0,1]^2$
1708: and we use the Hausdorff metric:
1709: \[
1710: d_H(P_1,P_2)=
1711: \sup_{x\in P_1}
1712: \inf_{y\in P_2}
1713: |x-y|
1714: +
1715: \sup_{x\in P_2}
1716: \inf_{y\in P_1}
1717: |x-y|.
1718: \]
1719:
1720:
1721:
1722:
1723: \begin{theorem}
1724: \label{paththeorem2}
1725: $P^{(n)*}\to P^*$ in distribution as $n\to\infty$.
1726: \end{theorem}
1727:
1728: \noindent\textit{Proof:}
1729: Choose a probability space
1730: on which the convergence in Theorem \ref{paththeorem}
1731: occurs almost surely. We will show that in this case
1732: $P^{(n)*}\to P^*$ a.s.\ also.
1733:
1734: First consider any point $y\in P^*$. We will show
1735: that for all sufficiently large $n$ there is a point
1736: of $P^{(n)*}$ within distance $\epsilon/2$ of $y$.
1737:
1738: There are two cases to consider.
1739:
1740: First, suppose that $y$ is a limit
1741: of some sequence of points $Y_i\in A^*$. Choose
1742: some $Y_i$ which is with distance $\epsilon/4$.
1743: For large enough $n$, we have $Y_i^{(n)}\in A^{(n)*}$
1744: and $|Y_i^{(n)}-Y_i|<\epsilon/4$ and we are done.
1745:
1746: Otherwise, $y$ is on a vertical or horizontal line
1747: between two points that \textit{are} limits
1748: of sequences $Y_i\in A^*$. Call these endpoints $y^-$ and $y^+$.
1749: For large enough $n$, there are $i^-$ and $i^+$ such that
1750: $Y_{i^-}, Y_{i^+}\in P^{(n)*}$ with
1751: $|Y_{i^-}^{(n)}-y^-|<\epsilon/2$ and
1752: $|Y_{i^+}^{(n)}-y^+|<\epsilon/2$, as above.
1753: Then by the increasing path property,
1754: the subpath of $P^{(n)*}$ joining $Y_{i^-}$ and $Y_{i^+}$
1755: passes within $\epsilon/2$ of every point on the
1756: line segment joining $y^-$ to $y^+$,
1757: and hence in particular within $\epsilon/2$ of $y$ as required.
1758:
1759: Now for some $m\in\NN$ set $\epsilon=1/m$
1760: and consider an increasing sequence of points
1761: $(0,0)=y^{(0)}, y^{(1)}, \dots, y^{(2m)}=(1,1)\in P^*$
1762: such that the $L_1$ distance $d_1(y^{(j)}, y^{(j+1)})$ between
1763: successive points is exactly $\epsilon$ for all $j$.
1764: (This is possible since $P^*$ is an increasing path and connected).
1765:
1766: Now for large enough $n$ there is a sequence
1767: $\ty^{(0)}, \ty^{(1)}, \dots, \ty^{(2m)}$
1768: of points of $P^{(n)*}$ such that
1769: $d_1(\ty^{(j)}, y^{(j)})<\epsilon/2$ for all $j$.
1770: Then necessarily $\ty^{(0)}, \ty^{(1)}, \dots, \ty^{(2m)}$
1771: is itself an increasing sequence
1772: and $d_1(\ty^{(j)}, \ty^{(j+1)})<2\epsilon$ for all $j$.
1773:
1774: Using the increasing path property for $P^*$,
1775: we have that every point of $P^*$ is within $L_1$
1776: distance $\epsilon/2$ of one of the $y^{(j)}$.
1777: Then since each $y^{(j)}$ is within $L_1$
1778: distance $\epsilon/2$ of a point of $P^{(n)*}$,
1779: we have that every point of $P^*$ is within $\epsilon$
1780: of $P^{(n)*}$.
1781:
1782: Similarly, the increasing path property for $P^{(n)*}$
1783: gives that every point of $P^{(n)*}$ is within
1784: $L_1$ distance $\epsilon$ of one of the $\ty^{(j)}$,
1785: and thus in turn within distance $3\epsilon/2$ of
1786: some point of $P^*$.
1787:
1788: Hence $d_H(P^*, P^{(n)*})<5\epsilon/2$,
1789: for all large enough $n$. This works for any $\epsilon=1/m$,
1790: so $P^{(n)*}\to P^*$ as required.\enpf
1791:
1792: \section{Stable process directed percolation}
1793: \label{stablesection}
1794:
1795: In this section we consider a directed last-passage
1796: percolation model based on stable L\'evy processes.
1797: This is the stable version of the Brownian
1798: directed percolation problem considered in \cite{OcoYor, hmo}.
1799:
1800: For $n\in\NN$, $t>0$, consider the random variable
1801: \[ L(n,t) = \sup_{0=t_0 \leq t_1 \leq...\leq t_n=t} \sum_{i=1}^n
1802: S^i_{t_{i-1}t_i}, \]
1803: where $S^i_{st} = S^i_t-S^i_s$, and $S^i$ are i.i.d.\ $\alpha$-stable
1804: processes for some $\alpha\in(0,2)$.
1805: The Brownian version of this problem, in which the
1806: stable processes are replaced by Brownian motions,
1807: gives a representation
1808: for the largest eigenvalue process in
1809: ``Hermitian Brownian motion''
1810: (a matrix-valued process whose marginal at any fixed time has the
1811: GUE distribution)
1812: and has been much studied in various contexts
1813: (see for example \cite{GUEs, GraTraWid, hmo, OcoYor}). We do
1814: not have a random matrix interpretation of this stable process
1815: version; however, an interesting connection could be
1816: to the case of Wigner random matrices with heavy-tailed
1817: entries considered by Soshnikov in \cite{Soshnikovheavy},
1818: where a scaling is obtained for the largest eigenvalues
1819: which corresponds to the one we have
1820: observed for the heavy-tailed last-passage percolation problem.
1821:
1822: We will show that the asymptotic behaviour of the distribution
1823: of $L(n,t)$, as $n$ becomes large, is again
1824: described by our continuous heavy-tailed last-passage
1825: directed percolation problem.
1826: Note that by scaling $L(n,tn) = t^{1/\alpha} L(n,n)$ in
1827: distribution and hence we can just consider $L(n,n)$.
1828:
1829: The processes $S^i$ have jump measure
1830: $c_+ x^{-\alpha-1}I_{x>0} +c_-|x|^{-\alpha-1}I_{x<0}$,
1831: for some $c_+>0$ and $c_-\geq 0$.
1832: The jumps play the role of weights for the percolation problem.
1833:
1834: \begin{theorem}
1835: \[
1836: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L(n,n)\to T
1837: \]
1838: in distribution as $n\to\infty$, where $T$ is the last-passage time
1839: in the continuous last-passage percolation model with index $\alpha$,
1840: defined at (\ref{Tdef}).
1841: \end{theorem}
1842:
1843: A short argument is available in the case $\alpha<1$
1844: (making use of the fact that the sum of all positive weights
1845: is finite), and we give this first.
1846:
1847: \subsection{Case $\alpha<1$}
1848: Let $M_1^{(n)}$, $M_2^{(n)}, \dots$
1849: be the set of positive jumps of the processes
1850: $S^1$, $S^2,\dots,S^n$ on the interval $[0,n]$,
1851: written in descending order.
1852:
1853: From the form of the jump measure, we can regard the ordered sequence
1854: of jumps as a Poisson random measure. Thus by a suitable
1855: transformation we can write the sequence of jumps in terms of a Poisson process and
1856: have that for any $n$,
1857: \begin{multline*}
1858: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}
1859: \Big(M_1^{(n)}, M_2^{(n)},\dots,M_k^{(n)},\dots\Big)
1860: \\
1861: \stackrel{d}{=}
1862: \Big(W_1^{-1/\alpha},(W_1+W_2)^{-1/\alpha},
1863: \dots,
1864: (W_1+\dots+W_k)^{-1/\alpha},
1865: \dots\Big),
1866: \end{multline*}
1867: where $W_i$ are i.i.d.\ exponential random variables with mean 1.
1868:
1869: Now let $L^{(n)+}_k$ be the maximal
1870: weight of a path, if one ignores all the weights except the $k$
1871: largest positive weights. Just as in the discrete case,
1872: one can show that
1873: \[
1874: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L_k^{(n)+}
1875: \to T_k
1876: \]
1877: in distribution as $n\to\infty$,
1878: and one also has $T_k\to T$ in distribution
1879: as $k\to\infty$,
1880: where $T_k$ and $T$ are the last passage times
1881: for the continuous problem as defined before.
1882:
1883: Now let $L^{(n)+}$ be the maximal weight
1884: of a path, if one considers all the positive
1885: weights but ignores all the negative ones. Then
1886: \[
1887: L^{(n)+} - L^{(n)+}_k
1888: \leq
1889: \sum_{r=k+1}^\infty M_r^{(n)}.
1890: \]
1891: Now the distribution of
1892: $n^{-2/\alpha}\sum_{r=k+1}^\infty M_r^{(n)}$
1893: does not depend on $n$, and converges to $0$
1894: in distribution as $k\to\infty$
1895: (since the sum of all the positive weights is a.s.\
1896: finite). So we have
1897: \[
1898: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L^{(n)+}
1899: \to T
1900: \]
1901: in distribution.
1902:
1903: Now consider the optimal path attaining $L^{(n)+}$. Consider the sum
1904: of (the absolute values of) all the negative weights along the path;
1905: call it $S^{(n)-}$. Since the positive and negative weights occur
1906: independently, $S^{(n)-}$ has just the same distribution as the sum of
1907: the negative weights for a single stable process between times $0$ and
1908: $n$. This is finite and on the scale $n^{1/\alpha}$ (in fact, the
1909: distribution of $n^{-1/\alpha}S^{(n)-}$ is independent of $n$). So
1910: certainly $n^{-2/\alpha}S^{(n)-}\to 0$ in distribution as
1911: $n\to\infty$. Since $L^{(n)+}-S^{(n)-}\leq L(n,n)\leq L^{(n)+}$, we
1912: obtain
1913: \[
1914: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L(n,n)
1915: \to T
1916: \]
1917: in distribution as $n\to\infty$, as required.
1918:
1919: \subsection{Case $1\leq \alpha<2$}
1920:
1921: \noindent\textit{Lower bound:}
1922:
1923: As before,
1924: \[
1925: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L_k^{(n)+}
1926: \to T_k
1927: \]
1928: in distribution as $n\to\infty$.
1929:
1930: Now consider a path realising $L_k^{(n)+}$ in this way;
1931: (for definiteness, say the first such path in the
1932: lexicographic order).
1933:
1934: Let $\tT^{(n)}_k$ be the total weight of this path,
1935: including all weights, and let
1936: $\tS=\tT^{(n)}_k-L_k^{(n)+}$.
1937:
1938: The distribution of $\tS$ can be described as follows.
1939: Generate the $n$ independent processes $S^1,\dots, S^n$
1940: from time $0$ to time $n$. Remove
1941: the $k$ largest positive jumps that occur in
1942: the $n$ processes in time $[0,n]$. Then $\tS$ has
1943: the distribution of the altered value of $S^1_n$,
1944: after the $k$ largest jumps from the set of processes
1945: have been removed.
1946:
1947: However, as $n\to\infty$, the probability
1948: that any of the $k$ largest jumps occur in the process
1949: $S_1$ tends to 0 (it is no larger than $k/n$); so with
1950: high probability, the procedure in the previous paragraph
1951: does not alter the value of $S^1_n$.
1952: Thus the limit in distribution of $n^{-1/\alpha}\tS$
1953: is the distribution of $n^{-1/\alpha}S^1_n$
1954: (which is independent of $n$). In particular,
1955: $n^{-2/\alpha}\tS\to 0$ in probability.
1956: Thus, for any $k$,
1957: \begin{align*}
1958: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}\tT^{(n)}_k
1959: &=
1960: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}\tS+
1961: \left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L_k^{(n)+}
1962: \\
1963: &\to T_k
1964: \end{align*}
1965: in distribution, as $n\to\infty$.
1966: Since $L(n,n)\geq \tT^{(n)}_k$ and $T_k\to T$ as $k\to\infty$,
1967: this establishes that $T$ is a lower bound
1968: for the limit in distribution of
1969: $\left(\frac{\alpha}{c_+}\right)^{1/\alpha}n^{-2/\alpha}L(n,n)$.
1970:
1971: \medskip
1972:
1973: \noindent\textit{Upper bound:}
1974:
1975: We first need a lemma on the tail behaviour
1976: of the difference between the supremum and infimum
1977: of the stable process on an interval:
1978:
1979: \begin{lemma}\label{jumpuplemma}
1980: Let $S$ be a stable process with index $\alpha$
1981: and jump measure
1982: $c_+ x^{-\alpha-1}I_{x>0} +c_-|x|^{-\alpha-1}I_{x<0}$,
1983: where $c_+>0$ and $c_-\geq 0$.
1984: Then
1985: \[
1986: \PP\Big(\sup_{0\leq s\leq t\leq 1}\{S_t-S_s\} > x\Big)
1987: \sim \frac{c_+}{\alpha} x^{-\alpha}
1988: \]
1989: as $x\to\infty$. That is,
1990: the quantity $\sup_{0\leq s\leq t\leq 1}\{S_t-S_s\}$
1991: has the same positive tail behaviour
1992: as the size of the largest positive jump of $S$
1993: in $[0,1]$
1994: (which is also the same as the upper tail of
1995: $S_1$ and of $\sup_{0\leq t\leq 1}S_t$).
1996: \end{lemma}
1997:
1998: \noindent\textit{Proof:}
1999: Let the running infimum and supremum processes be denoted by $I_t =
2000: \inf\{S_s:0\leq s\leq t\}$ and $S^*_t = \sup\{S_s: 0\leq s\leq t\}$
2001: respectively. Consider the reflected process $X_t = S_t - I_t$. Our
2002: aim is to determine the tail behaviour of $X^*_1 = \sup\{X_t:0\leq
2003: t\leq 1\}$.
2004: By Bertoin \cite{Bertoinbook}~Prop~VI.3,
2005: we know that for fixed $t$, the distribution of $X_t$
2006: is the same as that of $S^*_t$. By standard results
2007: (e.g. \cite{Bertoinbook}~Prop~VIII.4) we have
2008: \begin{equation}
2009: \PP(X_1>x) \sim \PP(S^*_1>x) \sim \frac{c_+}{\alpha} x^{-\alpha},
2010: \label{eq:tailasymp}
2011: \end{equation}
2012: as $x\to\infty$.
2013:
2014: Thus we just need to show that $X^*_1$ has the same tail as $X_1$.
2015: The proof is analogous to that of
2016: \cite{Bertoinbook}~Prop.~VIII.4,
2017: %% once we note
2018: %% that the reflected process is Markov by
2019: %% \cite{Bertoinbook}~Prop~VI.1, and that from the scaling for the stable
2020: %% process,
2021: %% $S_{\lambda t} \stackrel{d}{=} \lambda^{1/\alpha} S_t$,
2022: %% we have a scaling for the
2023: %% infimum and hence the reflected process itself has the scaling
2024: %% $X_{\lambda t} \stackrel{d}{=} \lambda^{1/\alpha} X_t$.
2025: %% \enpf
2026: and we reproduce the argument
2027: here. An easy consequence of (\ref{eq:tailasymp}) is that
2028: \[
2029: \liminf_{x\to\infty} \PP(X^*_1>x) x^\alpha \geq
2030: \frac{c_+}{\alpha}.
2031: \]
2032: Now fix $\epsilon>0$ and note that the reflected process is Markov by
2033: \cite{Bertoinbook}~Prop~VI.1. As the stable process scales in that
2034: $S_{\lambda t} \stackrel{d}{=} \lambda^{1/\alpha} S_t$,
2035: this property will be inherited by the
2036: infimum and hence the reflected process itself, giving
2037: $X_{\lambda t} \stackrel{d}{=} \lambda^{1/\alpha} X_t$.
2038: Since $S_t-X_t=I_t$ is decreasing in $t$, we also
2039: have that $X_1-X_t\geq S_1-S_t$ for all $t<1$.
2040: Applying these properties and denoting
2041: by $\tau_x$ the first hitting time of the interval $(x,\infty)$ by the
2042: reflected process $X$, we have
2043: \begin{align*}
2044: \PP\big(X_1>(1-\epsilon)x\big)
2045: &\geq
2046: \PP\big(X_1^*>x, X_1>(1-\epsilon)x\big)
2047: \\
2048: &\geq
2049: \int_0^1 \PP(\tau_x\in dt)\PP(X_1-X_t>-\epsilon x)
2050: \\
2051: &\geq
2052: \int_0^1 \PP(\tau_x\in dt)\PP(S_1-S_t>-\epsilon x)
2053: \\
2054: &=
2055: \int_0^1 \PP(\tau_x\in dt)\PP(S_{1-t}>-\epsilon x)
2056: \\
2057: &\geq
2058: \int_0^1 \PP(\tau_x\in dt)\PP(S_1>-\epsilon x)
2059: \\
2060: &=\PP(X_1^*>x)\PP(S_1>-\epsilon x).
2061: \end{align*}
2062: %
2063: %% \begin{eqnarray*}
2064: %% P(X_1>(1-\epsilon)x) &\geq & P(X^*_1>x,X_1>(1-\epsilon)x) \\
2065: %% &\geq & \int_0^1 P(X_{1-t}>-\epsilon x) P(\tau_x \in dt) \\
2066: %% &\geq & P(X^*_1>x) P(X_1>-\epsilon x).
2067: %% \end{eqnarray*}
2068: %
2069: As $\PP(S_1>-\epsilon x) \to 1$ as $x\to\infty$ we have that
2070: \[
2071: \limsup_{x\to\infty} P(X^*_1>x) x^{\alpha} \leq
2072: (1-\epsilon)^{-\alpha} c_+/\alpha,
2073: \]
2074: and, as $\epsilon$ is arbitrary, we have the result.\enpf
2075:
2076: Now for $1\leq i, j \leq n$, define
2077: \[
2078: X(i,j)=\sup_{j-1\leq s\leq t\leq j} \{S^i_t-S^i_s\}.
2079: \]
2080: Let $T(n,n)$ be the maximal passage time for the discrete
2081: model with weights $X(i,j)$.
2082: Then one can see by a direct sample path comparison that
2083: $L(n,n)\leq T(n,n)$.
2084:
2085: Applying Lemma \ref{jumpuplemma},
2086: $a_N\sim \left(\frac{\alpha}{c_+}\right)^{1/\alpha}N^{1/\alpha}$,
2087: where $a_N=\inf\{x:\PP(X(i,j)>x\leq 1/N\}$.
2088: Thus, applying Theorem \ref{Ttheorem} for the discrete model,
2089: \[
2090: \left(\frac{c_+}{\alpha}\right)^{1/\alpha}n^{-2/\alpha} T(n,n)
2091: \to T
2092: \]
2093: in distribution as $n\to\infty$.
2094: This gives the required upper bound in distribution
2095: for the limit of $L(n,n)$.
2096:
2097: \section{The case $\alpha=0$:
2098: convergence to the greedy path}
2099: \label{greedysection}
2100:
2101: In this section we consider the
2102: discrete last-passage percolation model in the
2103: case $\alpha=0$. The distribution
2104: $F$ is said to have a
2105: \textit{slowly varying tail}: for all $t>0$,
2106: \[
2107: \frac{1-F(tx)}{1-F(x)}\to 1 \textit{ as } x\to\infty;
2108: \]
2109: equivalently,
2110: for all $s<1$,
2111: \begin{equation}
2112: \label{first}
2113: \frac{F^{-1}(1-sv)}{F^{-1}(1-v)}\to \infty \text{ as } v\downarrow0.
2114: \end{equation}
2115:
2116: Now it is no longer possible to find a non-degenerate
2117: limit in distribution for $T^{(n)}$ as we did
2118: in Theorem \ref{Ttheorem}.
2119: Let $M^{(n)}_1$ be the maximum of an i.i.d.\ sample from $F$
2120: of size $n^2$ as before. Let $b_n$ be any sequence of constants.
2121: Then any limit point in distribution of the sequence
2122: $b_n^{-1} M^{(n)}_1$ must be concentrated on the set $\{0,\infty\}$.
2123: From Proposition \ref{greedyprop} below, the same is true
2124: if we replace $M^{(n)}_1$ by $T^{(n)}$.
2125:
2126: However, convergence in distribution of the optimal
2127: paths can still be obtained. In fact, the form
2128: of the limiting distribution has a particularly
2129: simple description in terms of a ``greedy algorithm''.
2130: We give a multifractal analysis of this limiting object
2131: in Section \ref{greedysubsection}.
2132:
2133: The limiting object is defined as follows.
2134: Given the locations $Y_1, Y_2,\dots$ i.i.d.\ uniform on $[0,1]^2$,
2135: let $\cC$, the set of increasing paths, be defined as at (\ref{cCdef}).
2136: We now define the
2137: \textit{greedy path} $A^*=A^*(Y_1, Y_2, \dots)\in \cC$
2138: recursively as follows.
2139: Let $1\in A^*$ always, and then,
2140: given $A^*\cap \{1,\dots,r\}$,
2141: let $r+1\in A^*$ if and only if $Y_{r+1}\sim Y_i$ for every
2142: $i\in A^*$, $i\leq r$.
2143: One can describe $A^*$ as the first member of $\cC$ in
2144: the lexicographic order.
2145:
2146: The discrete problem is defined in terms of the locations
2147: $(Y_i^{(n)})$ and weights $(M_i^{(n)})$ as before.
2148: Write $A^{(n)*}$ for the optimal path as in Section
2149: \ref{pathsection}. The following theorem
2150: gives the convergence of these optimal paths to the greedy path:
2151:
2152: \begin{theorem}
2153: \label{slowtheorem}
2154: $\Big(\big(Y^{(n)}_i\big)_{i\in\NN}, A^{(n)*}\Big)
2155: \to
2156: \Big(\big(Y_i\big)_{i\in\NN}, A^{*}\Big)$
2157: in distribution as $n\to\infty$.
2158: \end{theorem}
2159:
2160: Exactly as in Section \ref{paththeorem},
2161: one can also define $P^*$ and $P^{(n)*}$
2162: to represent the optimal paths regarded as subsets of $[0,1]^2$,
2163: and obtain the convergence in distribution of
2164: $P^{(n)*}$ to $P^*$ as $n\to\infty$ (under the Hausdorff metric).
2165: In fact, the situation is considerably simpler here;
2166: one can simply define $P^*$ to be the closure of
2167: $\bigcup_{i\in A^*} Y_i$, since by the results
2168: of Section \ref{greedysubsection} this set is connected w.p.\ 1.
2169:
2170: Theorem \ref{slowtheorem} will follow from the next proposition:
2171: \begin{proposition}
2172: \label{greedyprop}
2173: For all $r$,
2174: \[
2175: \PP\left(M^{(n)}_r > \sum_{i=r+1}^{n^2} M^{(n)}_i \right)\to 1
2176: \text{ as } n\to\infty.
2177: \]
2178: \end{proposition}
2179:
2180: \noindent\textbf{Proof of Theorem \ref{slowtheorem}}:
2181: Fix some $\epsilon>0$. As in Proposition \ref{coupledprop},
2182: for $n$ large enough,
2183: we can find a coupling such that, with probability at least $1-\epsilon$,
2184: \begin{gather}
2185: \nonumber
2186: \sum_{i=1}^k \left\| Y_i - Y^{(n)}_i \right\| < \epsilon
2187: \\
2188: \intertext{and}
2189: \label{thesame}
2190: \cC_k^{(n)}=\cC_k
2191: \end{gather}
2192: Suppose that (\ref{thesame}) holds and also that, for all $r\leq k$,
2193: $M^{(n)}_r > \sum_{i=r+1}^{n^2} M^{(n)}_i$.
2194: Then indeed
2195: \begin{equation}
2196: \label{thesame2}
2197: A^{(n)*}\intersect\{1,\dots,k\} = A^*\intersect\{1,\dots,k\},
2198: \end{equation}
2199: where $A^*$ is the ``greedy path''.
2200: Then using Proposition \ref{greedyprop},
2201: we can find $n$ such that
2202: (\ref{thesame2}) holds
2203: with probability at least $1-2\epsilon$.
2204: This gives the convergence in distribution in Theorem
2205: \ref{slowtheorem}.~$\hfill\Box$
2206:
2207: \subsection{Proof of Proposition \ref{greedyprop}}
2208:
2209: \begin{lemma}
2210: \label{theme}
2211: Let $\epsilon>0$ and $C>0$.
2212: There exists $U(C,\epsilon)<1$ such that if
2213: $U(C,\epsilon)<u_1<(1+\epsilon)u_2-\epsilon$, then
2214: \[
2215: F^{-1}(u_1)<F^{-1}(u_2)(1-u_1)^{-C}(1-u_2)^{C}.
2216: \]
2217: \end{lemma}
2218:
2219: \noindent\textit{Proof:}
2220: Let $t=(1+\epsilon)^{2C}$.
2221: From (\ref{first})
2222: there exists $V>0$ such that for all
2223: $v<V$,
2224: \[
2225: \frac{F^{-1}\left(1-\frac{v}{1+\epsilon}\right)}{F^{-1}(1-v)}>t.
2226: \]
2227: Iterating,
2228: \[
2229: \frac{F^{-1}\left(1-\frac{v}{(1+\epsilon)^m}\right)}{F^{-1}(1-v)}>t^m,
2230: \text{ for all }m\in\NN,
2231: \]
2232: and so in fact, for any $v_2<v_1<V$,
2233: \[
2234: \frac{F^{-1}(1-v_2)}{F^{-1}(1-v_1)}
2235: >t^{\left\lfloor
2236: \log_{1+\epsilon}\frac{v_1}{v_2}\right\rfloor}.
2237: \]
2238: Putting $u_1=1-v_1$, $u_2=1-v_2$ and $U(C,\epsilon)=1-V$,
2239: we have that for $U(C,\epsilon)<u_1<u_2$,
2240: \[
2241: F^{-1}(u_2)>
2242: t^{
2243: \left\lfloor
2244: \log_{1+\epsilon}\frac{1-u_1}{1-u_2}\right\rfloor}
2245: F^{-1}(u_1).
2246: \]
2247: Now whenever $z\geq 1$, then $\lfloor z\rfloor\geq z/2$,
2248: so restricting to $(1-u_1)>(1+\epsilon)(1-u_2)$
2249: we get
2250: \[
2251: F^{-1}(u_2) > t^{
2252: \frac{1}{2}\log_{1+\epsilon}\frac{1-u_1}{1-u_2}}
2253: F^{-1}(u_1),
2254: \]
2255: which rearranges to the desired result.$\hfill\Box$
2256:
2257: \begin{lemma}
2258: \label{variation}
2259: Fix $\epsilon>0$ and $C>0$.
2260: If $u_2\in(0,1)$ is sufficiently close to 1,
2261: then for all $u_1$ with
2262: $0<u_1<(1+\epsilon)u_2-\epsilon$,
2263: \[
2264: F^{-1}(u_1)<F^{-1}(u_2)(1-u_1)^{-C}(1-u_2)^{C}.
2265: \]
2266: \end{lemma}
2267:
2268: \noindent\textit{Proof:}
2269: From Lemma \ref{theme}, we already know that this is true
2270: when $u_1>U(C,\epsilon)$. Now if $u_1\leq U(C,\epsilon)$,
2271: then $F^{-1}(u_1)\leq F^{-1}(U(C,\epsilon))$. So it will suffice
2272: to show that for $u_2$ sufficiently close to 1, the RHS is always
2273: at least $F^{-1}(U(C,\epsilon))$. In fact, we will show
2274: that the RHS tends to $\infty$ as $u_2\uparrow 1$, uniformly in $u_1$.
2275:
2276: For any $u_2$, the RHS is minimised by $u_1=0$.
2277: So we wish to show that
2278: \begin{equation}
2279: \label{needed}
2280: F^{-1}(u_2)(1-u_2)^{C}
2281: \to\infty \text{ as } u_2\uparrow1.
2282: \end{equation}
2283:
2284: %To do this, we reuse Lemma \ref{theme} but replacing $t$ by some value
2285: %$\tildet>t$ and $u_1$ by some $\tu_1>V(\tildet,\epsilon)$.
2286:
2287: Fix $\tu_1>U(2C,\epsilon)$. Then Lemma \ref{theme} gives,
2288: for $u_2$ sufficiently close to 1,
2289: \begin{align*}
2290: F^{-1}(u_2)
2291: &\geq F^{-1}(\tu_1)
2292: (1-\tu_1)^{2C}(1-u_2)^{-2C}
2293: \\
2294: &=c(1-u_2)^{-2C}
2295: \end{align*}
2296: for some constant $c$. This gives the desired convergence to $\infty$
2297: in (\ref{needed}).~$\hfill\Box$
2298:
2299: \begin{lemma}
2300: \label{expectation}
2301: Let the r.v.\ $U$ be uniform on $(0,1)$.
2302: \[ \lim_{u\to1}\frac{\EE(F^{-1}(U)\given U\leq u)}{F^{-1}(u)(1-u)}
2303: =0. \]
2304: \end{lemma}
2305:
2306: \noindent\textit{Proof:}
2307: Take any $C>0$ and $\epsilon>0$.
2308: If $u$ is close enough to 1, then using Lemma \ref{variation},
2309: \begin{align*}
2310: \EE(F^{-1}(U); U\leq u) &= \int_0^u F^{-1}(v)dv \\
2311: &\leq \int_0^{(1+\epsilon)u-\epsilon} F^{-1}(u)(1-v)^{-C}(1-u)^{C}dv +
2312: \int_{(1+\epsilon)u-\epsilon}^u F^{-1}(u)dv \\
2313: &= F^{-1}(u)\left\{(1-u)^{C}\frac{(1+\epsilon)(1-u)^{-C+1}}{C-1} +
2314: \epsilon(1-u) \right\} \\
2315: &=F^{-1}(u)(1-u)\left\{\frac{(1+\epsilon)^{-C+1}}{C-1}+\epsilon \right\}.
2316: \end{align*}
2317:
2318: We can now choose $\epsilon$ as small as desired,
2319: and then $C$ as large as desired, to give an upper
2320: bound on
2321: \[ \limsup_{u\to1}\frac{\EE(F^{-1}(U); U\leq u)}{F^{-1}(u)(1-u)} \]
2322: which is arbitrarily close to 0.
2323:
2324: Finally, for $u\geq 1/2$,
2325: \[ \EE(F^{-1}(U)\given U\leq u) \leq 2\EE(F^{-1}(U);U\leq u) \]
2326: and the result follows.$\hfill\Box$
2327:
2328: \medskip
2329:
2330: \noindent\textbf{Proof of Proposition \ref{greedyprop}}:
2331: We use the representation
2332: \[
2333: \left(
2334: M^{(n)}_1,\dots,M^{(n)}_{n^2}
2335: \right)
2336: =
2337: \left(
2338: F^{-1}(U^{(n)}_1,\dots, U^{(n)}_{n^2})
2339: \right),
2340: \]
2341: where $(U^{(n)}_1,\dots, U^{(n)}_{n^2})$
2342: are the order statistics of an i.i.d.\ sample of size $n^2$ from
2343: the uniform distribution on $(0,1)$, written
2344: in decreasing order.
2345:
2346: We need to show that
2347: \[ \PP\left( \sum_{m=r+1}^{n^2} F^{-1}(U^{(n)}_m)\geq
2348: F^{-1}(U^{(n)}_r)\right)\to 0 \text{ as } n\to\infty. \]
2349: Let $R>0$ and suppose that $u>1-R/n^2$.
2350: Then
2351: \begin{align}
2352: \nonumber
2353: \frac{n^2\EE\big(F^{-1}(U)\given U\leq u\big)}{F^{-1}(u)}
2354: &=\frac{R \EE\big(F^{-1}(U)\given U\leq u\big)}{F^{-1}(u)R/n^2} \\
2355: \nonumber
2356: &\leq \frac{R\EE\big(F^{-1}(U)\given U\leq u\big)}{F^{-1}(u)(1-u)} \\
2357: \label{bound}
2358: &\leq R\sup_{u>1-R/n^2}\frac{\EE\big(F^{-1}(U)\given U\leq u\big)}
2359: {F^{-1}(u)(1-u)}.
2360: \end{align}
2361: Now for (almost) all $u$,
2362: \[
2363: \EE\left(\sum_{m=r+1}^{n^2} F^{-1}(U^{(n)}_m)
2364: \bigg\vert
2365: U^{(n)}_r=u\right)
2366: =(n^2-r)\EE\left(F^{-1}(U)\given U\leq u\right),
2367: \]
2368: so for (almost) all $u>1-R/n^2$ we have, from Markov's inequality,
2369: \begin{multline*}
2370: \PP\left(\sum_{m=r+1}^{n^2} F^{-1}(U^{(n)}_m)\geq F^{-1}(M^{(n)}_r)
2371: \bigg\vert U^{(n)}_r=u
2372: \right)
2373: \\
2374: \begin{aligned}
2375: &\leq
2376: \frac{1}{F^{-1}(u)}
2377: \EE\left(\sum_{m=r+1}^{n^2} F^{-1}(U^{(n)}_m)
2378: \bigg\vert
2379: U^{(n)}_r=u\right)
2380: \\
2381: &\leq
2382: n^2
2383: \frac{
2384: \EE\left(F^{-1}(U)\given U\leq u\right)
2385: }
2386: {F^{-1}(u)}
2387: \\
2388: &\leq
2389: R\sup_{u>1-R/n^2}
2390: \frac{
2391: \EE\big(F^{-1}(U)\given U\leq u\big)
2392: }
2393: {F^{-1}(u)(1-u)}
2394: \end{aligned}\end{multline*}
2395: using (\ref{bound}).
2396:
2397: So in fact, integrating over $u>1-R/n^2$,
2398: \begin{multline*}
2399: \PP\left(\sum_{m=r+1}^{n^2} F^{-1}(U^{(n)}_m)\geq F^{-1}(M^{(n)}_r)
2400: \bigg\vert U^{(n)}_r>1-R/n^2
2401: \right)
2402: \\
2403: \leq
2404: R\sup_{u>1-R/n^2}
2405: \frac{
2406: \EE\big(F^{-1}(U)\given U\leq u\big)
2407: }
2408: {F^{-1}(u)(1-u)}.
2409: \end{multline*}
2410: Then
2411: \begin{multline*}
2412: \limsup_{n\to\infty}
2413: \PP\left(\sum_{m=r+1}^{n^2} F^{-1}(U^{(n)}_m)\geq F^{-1}(U^{(n)}_r)\right)
2414: \\
2415: \leq
2416: \limsup_{n\to\infty}
2417: \PP\left(U^{(n)}_r\leq 1-R/n^2\right)
2418: +\limsup_{n\to\infty}
2419: R\sup_{u>1-R/n^2}
2420: \frac{
2421: \EE\big(F^{-1}(U)\given U\leq u\big)
2422: }
2423: {F^{-1}(u)(1-u)}.
2424: \end{multline*}
2425: The second term is 0 by Lemma \ref{expectation}, for every $R$.
2426: The first term is the probability $\PP(B_{n^2,R/n^2}\leq r-1)$,
2427: where $B_{n^2,R/n^2}$ is a binomial $(n^2,R/n^2)$ random variable.
2428: This converges as $n\to\infty$ to the probability
2429: $\PP(P_R\leq r-1)$ where $P_R$ is a Poisson mean $R$ random
2430: variable. Thus this probability can be made as small as desired by choosing $R$
2431: large, and hence the limsup is in fact $0$ as required.$\hfill\Box$
2432:
2433:
2434:
2435: \subsection{The properties of the greedy path}
2436: \label{greedysubsection}
2437:
2438: \begin{figure}[t]
2439: \centering
2440: \epsfig{figure=oo.ps, angle=270, width=1.3\linewidth}
2441: \caption{A simulation from the distribution of the ``greedy path''
2442: which occurs as the limit of the distribution of optimal paths
2443: in the case $\alpha=0$.
2444: \label{greedyfig}}
2445: \end{figure}
2446:
2447: In this section we discuss the properties of the greedy path
2448: which we have obtained as a distributional limit
2449: of the optimal path for the discrete problem when $\alpha=0$.
2450: The path can be regarded as a function $y=G(x)$ from $[0,1]$
2451: to itself in a natural way
2452: (specifically, one could define
2453: $G(x)=\sup\{y: Y_i=(x',y) \text{ for some } x'\leq x
2454: \text{ and some } i\in A^*\}$.
2455: See Figure \ref{greedyfig} for a realisation of the greedy path).
2456: This function $G$ is monotone non-decreasing, and hence defines
2457: a measure $\mu$ on $[0,1]$.
2458: We will show that $\mu$
2459: is a random self-similar measure which is singular with
2460: respect to Lebesgue measure, and we will be able to compute its
2461: multifractal spectrum.
2462:
2463: We recall the definition of the multifractal spectrum for our
2464: setting. Let $B_r(x)$
2465: denote the ball of radius $r$ around the point $x\in \br$.
2466: For $a\geq 0$,
2467: we define $E_a$, the set of
2468: points at which the measure has local dimension $a$ by
2469: \[ E_{a} = \left\{ x: \lim_{r\downarrow 0} \frac{\log
2470: \mu(B_r(x))}{\log r} = a \right\}. \]
2471: The multifractal spectrum is then defined to be
2472: \[ f(a) = \dim_H(E_{a}), \]
2473: where $\dim_H(A)$ denotes the Hausdorff dimension of a set $A$. In our
2474: setting we have
2475: \[ E_{a} = \left\{x\in [0,1]: \lim_{r\downarrow 0}
2476: \frac{\log(G(x+r)-G(x-r))}{\log r} = a \right\}. \]
2477:
2478: In particular we note that if $\mu$ had a density with respect to
2479: Lebesgue measure, then the spectrum would be the function
2480: $f(a)=0$ for all $a\neq 1$ and $f(1)=1$.
2481:
2482: There are a number of papers making rigorous the
2483: \textit{multifractal formalism}, the heuristic argument
2484: for computing the multifractal spectrum in terms of the Legendre
2485: transform of the moment measures, and we will be able to set our
2486: measure in a framework within which we can apply this formalism. The
2487: study of the multifractal spectrum for random self-similar measures is
2488: the topic of \cite{fal, ap, bhj} where the underlying assumptions are
2489: successively weakened.
2490:
2491: A \textit{scaling law} on a space $E$ consists of a probability space $(\Omega,
2492: \cf, \bp)$ and for each $\omega\in \Omega$ a collection of weights and
2493: maps $(\phi_1(\omega),p_1(\omega),\dots,\phi_N(\omega),p_N(\omega))$,
2494: where $p_i\in \br_+$ and $\phi_i:E \to E$ is a contraction with
2495: Lipschitz constant $r_i$. For a given scaling law a random
2496: self-similar measure is
2497: a measure $\mu$ which satisfies the distributional equality
2498: \[ \mu(\cdot) = \sum_{i=1}^N p_i \mu_i (\phi_i^{-1}(\cdot)), \]
2499: where $\mu_i$ are i.i.d.\ copies of $\mu$
2500: (independent of the weights and maps).
2501: The support of the measure is typically a random self-similar set.
2502:
2503: The multifractal formalism enables the
2504: multifractal spectrum for the random self-similar measure to be
2505: calculated in the following way. Let $m(q,\theta) = E( \sum_i p_i^q
2506: r_i^{\theta})$ and $\beta(q) = \inf\{\theta:m(q,\theta)\leq 1\}$.
2507: Under the formalism the multifractal spectrum is the Legendre transform of
2508: $\beta(q)$,
2509: \begin{equation}
2510: f(a) = \inf_{q\in\br}\{a q + \beta(q)\}. \label{eq:multform}
2511: \end{equation}
2512: We will now give a more formal version.
2513:
2514: We introduce a little notation.
2515: We write $\cT_n = \{1,\dots,N\}^n$ for
2516: the sequences which index the sets after $n$ applications of the scaling
2517: law, and $\cT$ for the tree
2518: $\cup_{n=0}^{\infty} \cT_n$.
2519: Let $(\Omega^{\otimes \cT},
2520: \cf^{\otimes \cT}, \bp^{\otimes\cT})$ denote the product probability space for
2521: random variables on the tree; for each node,
2522: we have an independent copy of the scaling law.
2523: Now for each $\bfi\in\cT_n$, define
2524: \begin{align*}
2525: p_{\bfi} =
2526: p_{i_1}(\omega_{\emptyset})p_{i_2}(\omega_{i_1})\dots
2527: p_{i_n}(\omega_{i_1 i_2\dots i_{n-1}})
2528: \\
2529: \intertext{and}
2530: r_{\bfi} =
2531: r_{i_1}(\omega_{\emptyset})r_{i_2}(\omega_{i_1})\dots
2532: r_{i_n}(\omega_{i_1 i_2\dots i_{n-1}}).
2533: \end{align*}
2534: The total mass of the random measure
2535: over the unit interval is given by
2536: $W = \lim_{n\toi} \sum_{\bfi\in\cT_{n}} p_{\bfi}$.
2537: In our setting we will consider random
2538: probability measures, so that $W=1$.
2539: The other limit random variable we need is
2540: is $W(q) =\lim_{n\to\infty} W_n(q)$ where
2541: $W_n(q) = \sum_{\bfi\in\cT_n} p_{\bfi}^q r_{\bfi}^{\beta}$.
2542:
2543: We also define the set $I_\beta\subseteq\RR$ as follows:
2544: $q*\in I_\beta$ if, for some $a\geq0$,
2545: the infimum $\inf_{q\in\RR}\{aq+\beta(q)\}$ is non-negative
2546: and is achieved at $q*$.
2547:
2548: Finally the \textit{strong open set condition}
2549: is that there is an open set $O$ such that,
2550: with probability 1, one has that
2551: $\phi_i(O)\cap\phi_j(O)=\emptyset$ for $i\neq j$, that $\cup_{i=1}^N
2552: \phi_i(O) \subset O$ and that $\mu(O)>0$.
2553:
2554: We now state a
2555: version of the main result of \cite{bhj} which can be applied in our
2556: setting.
2557:
2558: \begin{lemma}\label{lem:mf}
2559: Let $\mu$ be a random self-similar probability measure satisfying the
2560: strong open set condition. If the following three sets of
2561: conditions are satisfied:
2562: \begin{enumerate}
2563: \item $-\infty< \be \sum_i p_i\log p_i <0$;
2564: \item For all $q\in I_\beta$,
2565: $\be \sum_i (\log p_i) p_i^q r_i^{\beta(q)}<\infty$
2566: and
2567: $\be \sum_i (\log r_i) p_i^q r_i^{\beta(q)}<\infty$;
2568: \item
2569: For all $q\in I_\beta$,
2570: $\be
2571: \sum_i \left((\log p_i)^2 + (\log r_i)^2\right) p_i^q r_i^{\beta(q)}<\infty$
2572: and
2573: $\be W(q) \log_+W(q)<\infty$;
2574: \end{enumerate}
2575: then the multifractal formalism holds in the following sense.
2576: Define
2577: \[ \beta^*(a) = \inf_{q\in \br}\{ a q + \beta(q)\}. \]
2578: Then for any $a\geq0$,
2579: \[ f(a) = \max \{ \beta^*(a), 0\},\]
2580: with probability 1, and
2581: $E_a = \emptyset$ with probability 1 if
2582: $\beta^*(a)<0$.
2583: \end{lemma}
2584:
2585: We now return to the measure arising from our greedy path.
2586: From the definition, we can describe the greedy path
2587: recursively as follows: choose a point $Y=(Y(1), Y(2))$ uniformly in
2588: the box $[0,1]^2$. Then the original path
2589: is the union of two independent greedy paths,
2590: one scaled to lie in $[0,Y(1)]\times[0,Y(2)]$
2591: and the other to lie in $[Y(1),1]\times[Y(2),1]$.
2592: From this representation we can regard the induced measure
2593: as a random self-similar measure for a scaling law.
2594: Let $V$ and $\tV$ be independent uniform random variables.
2595: Then the scaling law has
2596: $N=2$ with the set of weights
2597: $(p_1,p_2)=(\tV,1-\tV)$ and the set of contractions $(\phi_1,\phi_2)$,
2598: where $\phi_1(x) = Vx$ has contraction factor $r_1=V$ and $\phi_2(x) =
2599: 1-(1-V)x$ has contraction factor $r_2=1-V$.
2600: Then the random self-similar measure $\mu$ satisfies
2601: $\mu(\cdot) = \sum_{i=1}^2 p_i \mu_i(\phi^{-1}_i(\cdot))$ in distribution.
2602: From the construction it is clear that the measure is a probability
2603: measure whose support is the unit interval.
2604:
2605: We note that our measure does not fit into the framework of \cite{fal}
2606: or \cite{ap}. The interval does not satisfy the strong separation
2607: condition as required in \cite{fal} (for a definition see \cite{fal})
2608: but instead it satisfies the strong open set condition with
2609: $O=(0,1)$. This ensures that the overlap between a pair of
2610: contractions applied to the unit interval occurs at one point.
2611: It also uses uniform random variables for the contraction ratios and
2612: therefore there is no strictly positive lower bound on the contraction
2613: ratios as required in \cite{ap}.
2614:
2615: \begin{theorem}\label{thm:mfs}
2616: The measure corresponding to the greedy path has for, any given $a\geq 0$,
2617: with probability 1,
2618: \[ f(a) = \left\{ \begin{array}{cl} \sqrt{8a} - a -1, &
2619: 3-2\sqrt{2} \leq a \leq 3+2\sqrt{2}, \\ 0, & \mbox{ otherwise.}
2620: \end{array} \right. \]
2621: If $a\in[0,3-2\sqrt{2})\cup(3+2\sqrt{2},\infty)$ then $E_a=\emptyset$
2622: with probability 1.
2623: \end{theorem}
2624:
2625: \noindent
2626: {\it Proof:}
2627: We determine the multifractal spectrum using the multifractal
2628: formalism. We need to consider $m(q,\theta) =
2629: E(\tV^q V^\theta + (1-\tV)^q (1-V)^\theta)$ and $\beta(q) =
2630: \inf\{\theta:m(q,\theta)\leq 1\}$. It is straightforward to compute
2631: these quantities and we have
2632: \[ m(q,\theta) = 2\int_0^1 y^q dy \int_0^1 x^\theta dx =
2633: \frac2{(1+q)(1+\theta)} \mbox{ for $q>-1, \theta>-1$}. \]
2634: Thus
2635: \[ \beta(q) = \frac2{q+1} -1, \]
2636: and we can calculate that
2637: \[
2638: \beta^*(a)=\sqrt{8a}-a-1
2639: \]
2640: for all $a\geq 0$.
2641: We have that $\beta^*(a)$ is non-negative on the interval
2642: $[3-2\sqrt2, 3+2\sqrt2]$ and negative elsewhere.
2643: Hence, if we can establish the three conditions of
2644: Lemma~\ref{lem:mf}, to justify the formalism, we will have proved our
2645: theorem.
2646:
2647: Another calculation gives that $I_\beta=[1-\sqrt2,1+\sqrt2]\subset(-1/2,3)$,
2648: and that for all $q\in I_\beta$, one also has $\beta(q)\in(-1/2,3)$.
2649:
2650: For condition (1), we can compute $\be \sum_i p_i\log p_i = -1/2$.
2651:
2652: For (2) straightforward calculations give $\be \sum_i (\log p_i) p_i^q
2653: r_i^{\beta} = 2((q+1)^2(1+\beta))^{-1}<\infty$ and $\be \sum_i (\log r_i) p_i^q
2654: r_i^{\beta} = 2((q+1)(1+\beta)^2)^{-1}<\infty$, for all $q>-1, \beta>-1$.
2655:
2656: Finally for the conditions (3) we have to do some work.
2657: It is easy to calculate the first condition
2658: \[ \be \sum_i \left((\log p_i)^2 + (\log r_i)^2\right) p_i^q
2659: r_i^{\beta} = \frac{2}{(1+q)^3(1+\beta)} +
2660: \frac{2}{(1+q)(1+\beta)^3}<\infty, \]
2661: for $q>-1,\beta>-1$.
2662:
2663: Thus we only have to verify the final condition.
2664:
2665: We begin by observing that all we need is to prove that for some
2666: $\epsilon>0$, $\be W(q)^{1+\epsilon}<\infty$ for each $q$. To do this
2667: we will show that $\be W_n(q)^{1+\epsilon}$ converges. First we
2668: observe that $W_n(q)$ is a martingale by the definition of $\beta(q)$
2669: and we compute the bracket process
2670: \begin{eqnarray*}
2671: [W(q)]_n &=&
2672: \sum_{i=1}^n \be \Big(\big(W_i(q)-W_{i-1}(q)\big)^2 | \cf_{i-1}\Big)
2673: \\
2674: &=& \sum_{i=1}^n \be \left(\left(\sum_{j\in \cT_{i-1}} p_j^qr_j^{\beta(q)}
2675: (\chi_j-1)\right)^2\Big| \cf_{i-1}\right),
2676: \end{eqnarray*}
2677: where $\chi_j = \tV_j^qV_j^{\beta(q)} + (1-\tV_j)^q(1-V_j)^{\beta(q)}$ and the
2678: $(V_j,\tV_j)$ are independent of each other and independent over $j$. By
2679: definition of $\beta$ we know that $\be \chi_j = 1$ and we can also
2680: compute
2681: \[ \be \chi_j^2 = \frac{2}{(1+2q)(1+2\beta)} +
2682: 2\frac{\Gamma(q+1)^2}{\Gamma(2q+2)}
2683: \frac{\Gamma(\beta+1)^2}{\Gamma(2\beta+2)}. \]
2684: One can easily check that this quantity is finite over
2685: $q$ and $\beta$ in $(-1/2,3)$, and hence for all $q\in I_\beta$.
2686: Thus we have
2687: \[ [W(q)]_n = \sum_{i=1}^n W_{i-1}(2q,2\beta) (\be \chi^2-1), \]
2688: where we write $W_n(a,b) = \sum_{\bfi\in\cT_n} p_{\bfi}^a r_{\bfi}^b$.
2689:
2690: With the bracket process we can control the moments of the martingale
2691: as for any $\gamma\geq 1$ there is a constant $c_{\gamma}$ such that
2692: \[ \be W_n(q)^\gamma \leq c_{\gamma} \be [W(q)]_n^{\gamma/2}. \]
2693: Thus we need to compute the moments of the bracket process. As we will take
2694: $1<\gamma<2$ and all terms in the sum are positive, straightforward estimates give
2695: \begin{eqnarray*}
2696: \be [W(q)]_n^{\gamma/2} &=& \be(\sum_{i=1}^n W_{i-1}(2q,2\beta) (\be
2697: \chi^2-1))^{\gamma/2} \\
2698: &\leq& \be \sum_{i=1}^n W_{i-1}(2q,2\beta)^{\gamma/2} (\be
2699: \chi^2-1)^{\gamma/2} \\
2700: &\leq& \sum_{i=1}^n \be W_{i-1}(\gamma q,\gamma\beta) (\be
2701: \chi^2-1)^{\gamma/2}.
2702: \end{eqnarray*}
2703: Thus we just need to find $\be W_i(\gamma q, \gamma\beta) = (\be(
2704: \tV^{\gamma q}V^{\gamma\beta} + (1-\tV)^{\gamma
2705: q}(1-V)^{\gamma\beta}))^i$. Integration gives
2706: \[ \be W_i(\gamma q, \gamma\beta) = \left(\frac{2}{(1+\gamma
2707: q)(1+\gamma\beta)}\right)^i = \left(\frac{2(1+q)}{(1+\gamma
2708: q)(1+q+\gamma(1-q))}\right)^i. \]
2709: Thus our result will hold if we can establish that over the range of
2710: $q$, we can find a $\gamma>1$ such that
2711: \[ \frac{2(1+q)}{(1+\gamma q)(1+q+\gamma(1-q))} < 1. \]
2712: A numerical calculation with the quadratic formula shows that this is
2713: the case and hence we have our result. \enpf
2714:
2715: The explicit form of the spectrum shows that $f(a)>0$ for
2716: $3-2\sqrt{2}<a<3+2\sqrt{2}$. We also observe that the set of
2717: points for which $a=2$ has full dimension 1. The next result
2718: shows that the measure corresponding to the path is ``singular'':
2719:
2720: \begin{corollary}
2721: With probability 1, the greedy path is continuous and strictly increasing,
2722: with zero derivative almost everywhere.
2723: \end{corollary}
2724:
2725: \noindent
2726: \emph{Proof:}
2727: These properties can be proved fairly directly from
2728: the construction of the greedy path, but here we
2729: deduce them immediately from the multifractal spectrum.
2730:
2731: By construction, the distribution of the path is symmetric in
2732: the $x$ and $y$ coordinates; hence the continuity property
2733: and the property that the path is strictly increasing are equivalent.
2734:
2735: Any point of discontinuity of the path belongs to the set $E_0$,
2736: by definition. But from Theorem \ref{thm:mfs} we have that
2737: $E_0$ is a.s.\ empty, so the path is a.s.\ continuous as desired.
2738:
2739: For the derivative, note that
2740: as $G$ is a distribution function
2741: it is almost everywhere differentiable with non-negative derivative.
2742: Let $D_1$ denote the set of points where the path is differentiable
2743: and has strictly positive derivative.
2744: It is straightforward to see that $D_1\subset E_1$.
2745: Thus as $f(1) = \sqrt{8}-2<1$, $\bp-a.s.$,
2746: we have that the Lebesgue measure of $D_1$ is 0 with probability 1. \enpf
2747:
2748: \section{Last-passage random fields and an Airy process}
2749: \label{Airysection}
2750:
2751: \begin{figure}[ht]
2752: \centering
2753: \epsfig{figure=1.0a.eps, width=0.85\linewidth}
2754: \caption{Figures \ref{Airyfig1}-\ref{Airyfig3}
2755: show simulations of the heavy-tailed Airy process $H_t$ for
2756: three different values of $\alpha$. Here $\alpha=1$.
2757: \label{Airyfig1}}
2758: \end{figure}
2759: \begin{figure}[ht]
2760: \centering
2761: \epsfig{figure=1.5a.eps, width=0.85\linewidth}
2762: \caption{The heavy-tailed
2763: Airy process $H_t$ in the case $\alpha=1.5$.\label{Airyfig2}}
2764: \end{figure}
2765: \begin{figure}[ht]
2766: \centering
2767: \epsfig{figure=1.99a.eps, width=0.85\linewidth}
2768: \caption{The heavy-tailed Airy process $H_t$
2769: in the case $\alpha=1.99$.\label{Airyfig3}}
2770: \end{figure}
2771:
2772: In this section we consider the extension of the results to the case
2773: of a random field. To do this we give a Poisson random measure
2774: construction of the continuous limit model.
2775:
2776: Let $\mu$ denote a Poisson random measure on $\br^3_+$ with intensity
2777: measure
2778: $\lambda(dx\,dy\,dz) = dx\,dy\,\alpha z^{-\alpha-1} dz$. That is, if
2779: $N(x,y,z) = \mu((0,x)\times(0,y)\times (z,\infty))= \int_0^x \int_0^y
2780: \int_z^{\infty} \mu(dx\,dy\,dz)$ denotes the
2781: number of points in $(0,x)\times(0,y)\times (z,\infty)$, then this has
2782: a Poisson distribution with mean $\be N(x,y,z) = xy z^{-\alpha}$,
2783: and the number of points in disjoint sets are independent.
2784:
2785: We can now extend our limiting model to this setting. To relate this
2786: to our original model we can consider the unit square in $\br_+^2$ and
2787: order the points of the Poisson random measure in decreasing order of their
2788: $z$-coordinates and we recover the sequence of weights
2789: $Z_i=M_i$ in the original model.
2790:
2791: We will write $(Y_i,Z_i)$ for a point of the Poisson random
2792: measure, where the points are labelled as above in that we regard
2793: $Y_i$ as the location in $\br^2_+$ and $Z_i$ as the weight in our continuous
2794: last passage percolation model. Let
2795: \[ \cC_{xy} = \{A \subset \bn
2796: \mbox{ such that for all $Y_i \sim Y_j$
2797: for all $i,j\in A$ and
2798: $Y_i \in [0,x]\times[0,y]$ for all $i\in A$}\} \]
2799: and define
2800: \[ T(x,y) = \sup_{A\in \cC_{xy}} \sum_{i\in A} Z_i. \]
2801: It is clear by Theorem~\ref{Ttheorem} that this random variable will
2802: exist for each fixed $x,y$. This can be extended
2803: to show that the random field
2804: $\{T(x,y),x>0,y>0\}$ exists and arises as the limit of the last
2805: passage model. Let
2806: \[ T^{(n)}(x,y) = \max_{\pi\in \Pi^n_{xy}} \sum_{v\in \pi} X(v), \]
2807: where $\Pi^n_{xy}$ is the set of directed
2808: paths from $(1,1)$ to $(\lceil nx \rceil, \lceil ny \rceil)$,
2809: and let
2810: \[ \tilde{T}^{(n)}(x,y) = \frac{T^{(n)}(x,y)}{a_{n^2}}. \]
2811:
2812: \begin{theorem}
2813: The field $\{T(x,y), x>0,y>0\}$
2814: exists almost surely and $\{\tilde{T}^{(n)}(x,y),
2815: x>0,y>0\} \to \{ T(x,y), x>0,y>0\}$ in the sense of convergence of
2816: finite dimensional distributions.
2817: \end{theorem}
2818:
2819: \noindent
2820: \emph{Proof:}
2821: Given a realisation of the PRM, the random variable
2822: $T(x,y)$ is well defined for all $x,y$, and over any finite box
2823: $[0,c_x]\times [0,c_y]$ we have $T(x,y)\leq T(c_x,c_y)$ whenever
2824: $x\leq c_x, y\leq c_y$; for any fixed $c_x,c_y$ we have
2825: $P(T(c_x,c_y)<\infty)=1$ and hence by countable additivity
2826: \[ P( T(x,y)<\infty \text{ for all } x,y ) =1. \]
2827:
2828: %For our PRM formulation of the problem we let
2829: %\[ T_{\epsilon}(x,y) = \sup_{A\in C^{\epsilon}_{xy}} \sum_{i\in A}
2830: %Z_i, \]
2831: %where
2832: %\[ C_{xy}^{\epsilon} = \{A \subset \bn \mbox{ such that for all
2833: % $i,j\in A, (X_i,Y_i) \sim (X_j,Y_j)$ with
2834: % $(X_i,Y_i) \in [0,x]\times[0,y]$ and $Z_i\geq \epsilon$}\}. \]
2835:
2836: %The existence of the field follows from the result that for any finite
2837: %region $(0,c_x)\times(0,c_y)$ we have $\lim_{\epsilon\to 0}
2838: %T_{\epsilon}(c_x,c_y) = T(c_x,x_y)<\infty$ and hence for any finite
2839: %set of points $(x_1,y_1),\dots,(x_n,y_n)$ with $(x_i,y_i) \leq
2840: %(c_x,c_y)$ for $i=1,\dots,n$ we have
2841: %\[ T_{\epsilon}(x_1,y_1), \dots, T_{\epsilon}(x_n,y_n) \leq
2842: %T_{\epsilon}(c_x,c_y) \leq T(c_x,x_y). \]
2843: %The $T_{\epsilon}$ are monotone increasing and bounded above and this
2844: %allows us to pass to the limit. Hence
2845: %\[ \lim_{\epsilon\to 0} T_{\epsilon}(x,y) = T(x,y), \;\;\forall x,y
2846: %\in (0,c_x)\times (0,c_y). \]
2847: %Thus for any finite set of points we have the joint convergence.
2848:
2849: We now need to establish the convergence of finite dimensional distributions.
2850: For one particular point we already have the one-dimensional
2851: convergence result given in Theorem~\ref{Ttheorem}.
2852: Exactly the same couplings
2853: between the discrete and continuous problems that we
2854: used to prove Theorem~\ref{Ttheorem}
2855: can be applied to extend the result to a finite collection of points
2856: within any finite box. The size of the box is
2857: arbitrary and we obtain convergence of all the finite dimensional
2858: distributions. \enpf
2859:
2860: We now proceed to define a stationary field on the whole of
2861: $\br^2$. To do this we observe that by simple scaling $N(\lambda x, \nu y,
2862: (\lambda\nu)^{1/\alpha} z) = N(x,y,z)$ in distribution. In particular
2863: we have
2864: \begin{equation}
2865: T(\lambda x, \nu y) = (\lambda\nu)^{1/\alpha} T(x,y) \mbox{ in
2866: distribution.}\label{eq:scale}
2867: \end{equation}
2868: Now put
2869: \[ \Theta(u_1,u_2) = \exp\left(-\frac{(u_1+u_2)}{\alpha}\right)
2870: T(e^{u_1},e^{u_2})
2871: \]
2872: for all $u_1,u_2\in\RR$.
2873: Then we have that for any $v_1,v_2\in\RR$,
2874: the collections
2875: $\{\Theta(u_1,u_2), u_1,u_2\in\RR\}$
2876: and
2877: $\{\Theta(u_1+v_1,u_2+v_2), u_1,u_2\in\RR\}$
2878: have the same distribution.
2879:
2880: The next two results concern the moments and correlations of
2881: this stationary field.
2882:
2883: \begin{proposition}\label{prop:fieldmoment}
2884: For all $\beta\in(0,\alpha)$ and all $\bfu\in\br^2$, we have
2885: $\be \Theta(\bfu)^{\beta} = \be T(1,1)^{\beta}<\infty$.
2886: \end{proposition}
2887:
2888: \noindent
2889: \emph{Proof:} This is a continuation of the argument given to
2890: establish the existence of $T(1,1)$ in the proof of
2891: Lemma~\ref{Slemma}. Recalling the setting of that proof,
2892: we have $T(1,1)=S_0\leq U_0$,
2893: where for $k\geq 0$, we defined
2894: $U_k=\sum_{i=k+1}^\infty L_i (M_i-M_{i+1})$.
2895: Here $L_i$ is the largest number of the first $i$ locations
2896: that can be included in an increasing path;
2897: there is a constant $c$ such that
2898: $\EE L_i\leq c\sqrt i$ and $\EE L_i^2\leq ci$ for all $i$,
2899: and the collections $(M_i)$ and $(L_i)$ are independent.
2900:
2901: Since $T(1,1)\leq U_0$, it will be enough to show that $\EE U_0^\beta<\infty$.
2902: If a finite collection of random variables each have a
2903: finite $\beta$th moment, then so does their sum.
2904: Hence, since $U_0=L_1(M_1-M_2)+\dots+L_k(M_k-M_{k+1}) +U_k$,
2905: it's enough to show both of the following:
2906: \begin{itemize}
2907: \item[(i)]for any $i$, $\EE\big[L_i(M_i-M_{i+1})\big]^\beta<\infty$;
2908: \item[(ii)]for some $k$, $\EE U_k^\beta<\infty$.
2909: \end{itemize}
2910: For property (i), note that $\big[L_i(M_i-M_{i+1})\big]^\beta<(iM_1)^\beta$,
2911: so it's enough to show that $\EE M_1^\beta<\infty$.
2912: But $M_1=W_1^{-1/\alpha}$, where $W_1$ has exponential distribution
2913: with mean 1. Thus $\EE M_1^\beta=\int_0^\infty w^{-\beta/\alpha}e^{-w}dw$,
2914: which is finite for all $\beta<\alpha$ as required.
2915:
2916: So it remains to show (ii).
2917: Since $\beta<\alpha<2$, it is enough to show that $\EE U_k^2<\infty$
2918: for all large enough $k$, and this is what we will do.
2919:
2920: Recall that we can write $M_i=(W_1+\dots+W_i)^{-1/\alpha}$
2921: where $W_j$ are i.i.d. exponential random variables
2922: with mean 1. We also write $V_i=W_1+\dots+W_i$, so that $M_i=V_i^{-1/\alpha}$.
2923: Then, using the fact that $(1+x)^{-1/\alpha}>1-x/\alpha$ for all $x>0$,
2924: we have that for all $i$,
2925: \begin{align}
2926: \nonumber
2927: \EE(M_i-M_{i+1})^2
2928: &=
2929: \EE\left[
2930: V_i^{-1/\alpha}\left(
2931: 1-\left(\frac{V_{i+1}}{V_i}\right)^{-1/\alpha}
2932: \right)
2933: \right]^2
2934: \\
2935: \nonumber
2936: &=
2937: \EE\left[
2938: V_i^{-2/\alpha}\left(
2939: 1-\left(1+\frac{W_{i+1}}{V_i}\right)^{-1/\alpha}
2940: \right)^2
2941: \right]
2942: \\
2943: \nonumber
2944: &\leq
2945: \EE\left[
2946: \frac{1}{\alpha^2}V_i^{-2/\alpha}V_i^{-2}W_{i+1}^2
2947: \right]
2948: \\
2949: \nonumber
2950: &=C\EE V_i^{-2-2/\alpha},
2951: \end{align}
2952: for some constant $C$, since $V_i$ and $W_{i+1}$ are independent.
2953: Arguing as at (\ref{Gammacalc}) and (\ref{Gammafact}),
2954: we have that $V_i$ has Gamma$(i,1)$ distribution, and we
2955: obtain that for some constant $\tC$ and all large enough $i$
2956: \begin{equation}
2957: \label{Mbound}
2958: \EE(M_i-M_{i+1})^2\leq \tC i^{-2-2/\alpha}.
2959: \end{equation}
2960:
2961: Suppose $k$ is large enough that (\ref{Mbound}) holds for all $i>k$.
2962: Then using Cauchy-Schwarz, we obtain that for all $\epsilon>0$,
2963: \begin{align*}
2964: U_k
2965: &=\sum_{i=k+1}^\infty L_i(M_i-M_{i+1})
2966: \\
2967: &\leq \left[
2968: \sum_{i=k+1}^\infty
2969: \left(
2970: i^{-1/2-\epsilon}
2971: \right)^2
2972: \right]^{1/2}
2973: \left[
2974: \sum_{i=k+1}^\infty
2975: \left(
2976: i^{1/2+\epsilon}
2977: L_i
2978: (M_i-M_{i+1})
2979: \right)^2
2980: \right]^{1/2}.
2981: \end{align*}
2982: The first sum is finite for any $\epsilon>0$.
2983: Thus squaring and taking expectations, we have that
2984: \begin{align*}
2985: \EE U_k^2
2986: &\leq
2987: c'\sum_{i=k+1}^\infty
2988: i^{1+\epsilon}\EE L_i^2 \EE(M_i-M_{i+1})^2
2989: \\
2990: &\leq
2991: c''\sum_{i=k+1}^\infty i^{1+\epsilon} i i^{-2-2/\alpha}
2992: \\
2993: &=c''\sum_{i=k+1}^\infty i^{-2/\alpha+\epsilon},
2994: \end{align*}
2995: for some constants $c', c''$. Since $\alpha<2$,
2996: this sum is finite for small enough $\epsilon$,
2997: and so $\EE U_k^2<\infty$ as desired.\enpf
2998:
2999: \begin{proposition}
3000: \label{prop:holder}
3001: For all $\beta<\alpha$ and all $\bfu, \bfv\in\RR_+^2$,
3002: we have
3003: \[
3004: \be |T(\bfu)-T(\bfv)|^{\beta} \leq
3005: 4\max\{\|\bfu\|,\|\bfv\|\}^{\beta/\alpha} \|\bfu-\bfv\|^{\beta/\alpha}
3006: \be T({\bf 1})^{\beta}.
3007: \]
3008: \end{proposition}
3009:
3010: \noindent
3011: \emph{Proof:} Recall that $T(\bfu)$ is increasing in the partial order
3012: on $\br_+^2$. We just need to consider the two cases where $\bfu \leq
3013: \bfv$ and where they are not comparable, so that, say, $u_1<v_1$ and
3014: $u_2>v_2$.
3015:
3016: For $\bfu\leq\bfv$, it is a simple observation that
3017: \begin{equation}
3018: \label{tvtubound}
3019: T(\bfv) -T(\bfu) \leq T\big((0,u_2),\bfv\big) + T\big((u_1,0),\bfv\big),
3020: \end{equation}
3021: where, for $\bx\leq\by\in\RR_+^2$,
3022: $T(\bx,\by)$ denotes the maximal weight of an increasing path
3023: from $\bx$ to $\by$.
3024: By the scaling in the field we have the distributional relationships
3025: \begin{equation}\label{scalingrel}
3026: T\big((a,b),(c,d)\big)
3027: \stackrel{d}{=}
3028: T(c-a,d-b)
3029: \stackrel{d}{=}
3030: (c-a)^{1/\alpha}(d-b)^{1/\alpha}T(1,1).
3031: \end{equation}
3032: For any random variables $X_1$, $X_2$, we have
3033: $\EE(X_1+X_2)^\beta\leq 2^{\max\{\beta,1\}}
3034: \max\{\EE X_1^\beta, \EE X_2^\beta\}$.
3035: Thus from (\ref{tvtubound}) and (\ref{scalingrel}) we obtain
3036: \begin{align}
3037: \nonumber
3038: \be |T(\bfv)-T(\bfu)|^{\beta}
3039: &\leq 2^{\max\{\beta,1\}}
3040: \max\left\{
3041: \EE T(v_1-u_1,v_2)^\beta,
3042: \EE T(v_1,v_2-u_2)^\beta
3043: \right\}
3044: \\
3045: \label{firstform}
3046: &\leq
3047: 4 \EE T({\bf 1})^\beta \max\left\{
3048: (v_1-u_1)^{\beta/\alpha}v_2^{\beta/\alpha},
3049: v_1^{\beta/\alpha}(v_2-u_2)^{\beta/\alpha}
3050: \right\}.
3051: \end{align}
3052: For the case $u_1<v_1$, $v_2<u_2$, we observe similarly that
3053: \begin{equation}
3054: \label{eq:tuv3}
3055: |T(\bv)-T(\bu)|\leq T\big((u_1,0), (v_1,u_2)\big)+T\big((0,v_2),(v_1,u_2)\big).
3056: \end{equation}
3057: Raising to the power $\beta$ and taking expectations as above,
3058: we obtain that
3059: \begin{equation}\label{secondform}
3060: \EE |T(\bv)-T(\bu)|^\beta
3061: \leq 4\EE T({\bf 1})^{\beta}\max\left\{
3062: (v_1-u_1)^{\beta/\alpha}v_2^{\beta/\alpha},
3063: (u_2-v_2)^{\beta/\alpha}u_1^{\beta/\alpha}
3064: \right\}.
3065: \end{equation}
3066: Combining the estimates from
3067: (\ref{firstform}) and (\ref{secondform}) now gives the result.
3068: \enpf
3069:
3070: We are now ready to define our analogue of the Airy process. If we set
3071: $H_u = \Theta(u,-u) = T(e^u,e^{-u})$, we have a one-dimensional stationary
3072: process $\{H_u,u\in\br\}$,
3073: as the processes $(H_{u+t}, u\in\br)$ and $(H_u, u\in\br)$
3074: have the same distribution for all
3075: $t\in \br$. We note that as $H_0 = T(1,1)$, the marginal
3076: distribution for our stationary process is the distribution of the
3077: limit random variable in our continuous model for heavy-tailed last
3078: passage percolation. By applying the estimates for the random field we
3079: have estimates on the H\"older continuity of the heavy-tailed Airy
3080: process.
3081:
3082: \begin{corollary}
3083: For $\beta<\alpha$, we have:
3084: \begin{itemize}
3085: \item[(i)] $\be H_0^\beta <\infty$.
3086: \item[(ii)] For each $\tau>0$ and all $u,v\in[-\tau,\tau]$,
3087: \[ \be |H_u-H_v|^{\beta} \leq 2e^{2\beta\tau/\alpha}
3088: |u-v|^{\beta/\alpha} \be H_0^\beta. \]
3089: \end{itemize}
3090: \end{corollary}
3091:
3092: \noindent
3093: \emph{Proof:} The first part is
3094: Proposition~\ref{prop:fieldmoment}. The second follows from the second
3095: part of the proof of Proposition~\ref{prop:holder} as, assuming $u<v$
3096: and using (\ref{eq:tuv3}) with $v_1=e^v, v_2=e^{-v}, u_1=e^u,u_2=e^{-u}$,
3097: \begin{eqnarray*}
3098: \be |H_u-H_v|^{\beta} &\leq& (e^v(e^{-u}-e^{-v}))^{\beta/\alpha} \be
3099: H_0^{\beta} + ((e^v-e^u)e^{-u})^{\beta/\alpha} \be H_0^{\beta} \\
3100: &=& 2 (e^{v-u}-1)^{\beta/\alpha} \be H_0^{\beta} \\
3101: &\leq & 2(v-u)^{\beta/\alpha} e^{(v-u)\beta/\alpha}\be H_0^{\beta},
3102: \end{eqnarray*}
3103: and the result follows. \enpf
3104:
3105: Finally we give a weak convergence result for our
3106: heavy-tailed Airy process.
3107:
3108: \begin{theorem}
3109: The sequence $\{a_{n^2}^{-1} T^{(n)}(e^u,e^{-u})\}_{u\in [-\tau,\tau]}$
3110: converges weakly to $\{H_u\}_{u\in [-\tau,\tau]}$ in $D[-\tau,\tau]$.
3111: \end{theorem}
3112:
3113: \noindent
3114: \emph{Proof:}
3115: We follow the approach outlined in \cite{Resnick86} for such
3116: weak convergence problems, in particular the proof of
3117: \cite{Resnick86}~Proposition~3.4.
3118:
3119: Since we restrict to $u\in[-\tau, \tau]$,
3120: we can consider only those points of the
3121: PRM whose locations fall in the box $[0,c_x]\times[0,c_y]$
3122: where $c_x=c_y=e^\tau$. Thus we can write
3123: the points as a sequence $(Y_i, Z_i), i=1,2,\dots$
3124: such that the sequence $(Z_i)$ is decreasing and such that
3125: $Y_i\in [0,c_x]\times [0,c_y]$ for all $i$.
3126:
3127: Define $H^{k}_u = T_k(e^u,e^{-u})$ where
3128: \[ T_k(x,y) = \sup_{A\in \cC^k_{xy}} \sum_{i\in A} Z_i, \]
3129: with
3130: \[ \cC^k_{xy} = \{A \subset\{1,\dots,k\} : Y_i\sim Y_j \,\forall
3131: i,j\in A\}. \]
3132:
3133: As before, we need a corresponding formulation of the
3134: discrete model. For given $n$,
3135: we work on the box $B^n=\{1,\dots,\lceil ne^\tau\rceil\}^2$.
3136: The weight at a point $y\in B^n$ has distribution $F$,
3137: independently for different points.
3138: We represent the weights and their positions
3139: by a vector $(Y^n_i, M^n_i,1\leq i\leq \lceil ne^\tau\rceil^2)$,
3140: where $M^n_i$ form a decreasing sequence.
3141: The interpretation is that $Y^n_i$ is the location
3142: of the $i$th largest weight $M^n_i$ in the box $B^n$.
3143:
3144: We can then define
3145: $H^{(n),k} =
3146: \tilde{T}^{(n)}_k(\lceil ne^u\rceil ,\lceil ne^{-u}\rceil)$
3147: where
3148: \[ \tilde{T}^{(n)}_k(\lceil nx \rceil ,\lceil ny \rceil)
3149: = \sup_{A\in \cC^{(n),k}_{xy}} \sum_{i\in A}
3150: a_{n^2}^{-1} M^{(n)}_i, \]
3151: with
3152: \[ \cC^{(n),k}_{xy} = \{A \subset\{1,\dots,k\wedge n^2\} :
3153: \forall i,j\in A, Y^{(n)}_i \sim Y^{(n)}_j\}. \]
3154:
3155: For the proof we will establish that $H^{(n),k}$ converges weakly
3156: to $H^k$, that $H^k \to H$ locally uniformly and that
3157: \[ \lim_{k\to\infty} \limsup_{n\to\infty}
3158: \bp(\rho(H^{(n),k},H^{(n)})>\epsilon) = 0, \]
3159: for each $\epsilon>0$, where $\rho$ is the metric for the Skorohod
3160: topology.
3161:
3162: % \[ \rho(H^{(n)},H) \leq \rho(H^{(n)},H^{(n),\delta}) +
3163: %\rho(H^{(n),\delta},H^{\delta}) +
3164: %\rho(H^{\delta},H), \]
3165:
3166: We begin by establishing $H^k \to H$ as $k\to\infty$. This is
3167: a consequence of the construction via a PRM. For each $u$, $H^k_u$ is an
3168: increasing function of $k$ and converges to $H_u$.
3169: With probability 1, this holds
3170: uniformly for each $u\in[-\tau, \tau]$,
3171: since for all such $u$ we have
3172: \[
3173: 0\leq H_u-H_u^k \leq
3174: \sup_{A\in \cC_{e^\tau, e^\tau}}\sum_{i\in A, i>k} M_i;
3175: \]
3176: the upper bound is finite with probability 1 exactly as in
3177: Lemma \ref{Slemma}.
3178:
3179: Next we wish to show the
3180: weak convergence of $H^{(n),k}$ to $H^k$.
3181: This can be done by an extension of the method
3182: of Proposition \ref{coupledprop}.
3183: Note that with probability 1,
3184: no two points of the PRM share a vertical coordinate
3185: or a horizontal coordinate, and in addition
3186: no point of the PRM falls on the line
3187: parametrised by $(x,y)=(e^u, e^{-u})$.
3188: Then under the same couplings used in the proof
3189: of Proposition \ref{coupledprop},
3190: one obtains that with probability 1,
3191: $H^{(n),k}\to H^k$ in the Skorohod space.
3192: Thus this weak convergence also holds as desired.
3193:
3194: Finally we need to control $\rho(H^{(n),k},H^{(n)})$.
3195: Fix an $\epsilon>0$.
3196: Consider
3197: \begin{eqnarray*}
3198: \bp\left(\rho(H^{(n),k},H^{(n)}) > \epsilon\right)
3199: &\leq &
3200: \bp\left(\sup_{-\tau<u<\tau}
3201: \left|H^{(n),k}_u - H^{(n)}_u\right|>\epsilon\right)
3202: \\
3203: &=&
3204: \bp\left(\sup_{-\tau<u<\tau} \left|\tilde{T}_k^{(n)}(e^u,e^{-u}) -
3205: \tilde{T}^{(n)}(e^u,e^{-u})\right|>\epsilon\right)
3206: \\
3207: &=&
3208: \bp\left(\sup_{-\tau<u<\tau}
3209: \left|
3210: \sup_{A\in C^{(n),k}_{e^u,e^{-u}}} \sum_{i\in A}
3211: a_{n^2}^{-1} M_i^{n} - \sup_{A\in C^{(n)}_{e^u,e^{-u}}} \sum_{i\in A}
3212: a_{n^2}^{-1} M_i^{n}
3213: \right|
3214: >\epsilon\right)
3215: \\
3216: &\leq&
3217: \bp\left(\sup_{-\tau<u<\tau}
3218: \left|
3219: \tilde{S}^{(n)}_k(e^u,e^{-u})
3220: \right|
3221: >\epsilon\right),
3222: \end{eqnarray*}
3223: where $\tilde{S}^{(n)}_k(e^u,e^{-u}) = \sup_{A\in
3224: C^{(n)}_{e^u,e^{-u}}} \sum_{i\in A, i>k} a_{n^2}^{-1} M_i^{n}$. By
3225: monotonicity we have
3226: \[ \sup_{-\tau<u<\tau} \left| \tilde{S}^{(n)}_k(e^u,e^{-u})\right| \leq
3227: \tilde{S}^{(n)}_k(e^\tau,e^\tau), \]
3228: and using the scaling
3229: \[ \bp\Big(\rho(H^{(n),k},H^{(n)}\Big) > \epsilon) \leq
3230: \bp\Big(\tilde{S}^{(n)}_{k}(1,1) > \epsilon e^{-2\tau/\alpha}\Big). \]
3231:
3232: By Proposition~\ref{Aprop} we have that
3233: $\bp\big(\tilde{S}^{(n)}_k > \epsilon e^{-2\tau/\alpha}\big) \to 0$
3234: as $k\to \infty$ uniformly in $n$.
3235: Thus indeed we have
3236: \[ \lim_{k \to\infty} \limsup_{n\to \infty}
3237: P\big(\rho(H^{(n),k},H^{(n)}\big) >\epsilon) =0. \]
3238: Putting the three pieces together we have shown the weak convergence. \enpf
3239:
3240: \section{Higher-dimensional heavy-tailed last passage percolation}
3241: \label{dsection}
3242: Up to this point we have considered only two-dimensional models.
3243: In this section we indicate how to extend most of the results
3244: to higher dimensions in a natural way.
3245:
3246: For general $d\geq 2$, we consider the passage time
3247: from the point $(1,1,\dots,1)$ to the point $(n,n,\dots, n)$.
3248:
3249: We now consider a sequence of locations $Y_i^{(n)}$,
3250: $i=1,2,\dots, n^d$ which form a uniform random
3251: permutation of the set $\{1/n, 2/n, \dots, 1\}^d\subset [0,1]^d$,
3252: and a corresponding sequence of weights $M_i^{(n)}$, $i=1,2,\dots,n^d$
3253: which are given by the order statistics, in decreasing order,
3254: of a sample of size $n^d$ from the underlying weight distribution $F$.
3255: We now assume that the tail of $F$ is regularly varying
3256: with index $\alpha<d$.
3257:
3258: Defining $\cC^{(n)}$ as before, we set
3259: $T^{(n)}=\sup_{A\in\cC^{(n)}} \sum_{i\in A} M_i^{(n)}$,
3260: and $\tT^{(n)}=a_{n^d}^{-1} T^{(n)}$.
3261:
3262: The continuous model is defined just as before;
3263: the locations $Y_i$ are now drawn i.i.d.\ and uniformly at random
3264: from the box $[0,1]^d$ rather than the square $[0,1]^2$.
3265:
3266: Then $\tT^{(n)}\to T$ in distribution as $n\to\infty$;
3267: the method of proof is essentially identical to that
3268: used for Theorem \ref{Ttheorem} in the case $d=2$.
3269:
3270: The multivariate extensions described in Sections
3271: \ref{preAiry} and \ref{Airysection}
3272: go through in an analogous way.
3273: For example, we can now obtain a process $\Theta$ which is
3274: stationary on $\RR^d$ such that
3275: \[
3276: \Bigg\{ \exp\left(-\frac{u_1+\dots+u_d}{\alpha}\right)
3277: a_{n^d}^{-1}T^{(n)}\left(e^{u_1},\dots,e^{u_d}\right), \bu\in\RR^d\Bigg\}
3278: \to \Big\{\Theta(u_1,\dots,u_d), \bu\in\RR^d\Big\}
3279: \]
3280: as $n\to\infty$, in the sense of convergence of
3281: finite-dimensional distributions;
3282: here $T^{(n)}(u_1,\dots,u_d)$ is the
3283: maximal weight of a path from $(1,\dots,1)$ to
3284: the point $(\lceil n u_1 \rceil,
3285: \dots, \lceil n u_d \rceil)$.
3286:
3287: We turn to the path convergence as developed in Section
3288: \ref{pathsection}.
3289: Proposition \ref{uniquenessprop} and Theorem \ref{paththeorem}
3290: extend easily, with the same method of proof.
3291: However, extending Theorem \ref{paththeorem2},
3292: concerning the convergence of optimal paths
3293: viewed as random subsets of $[0,1]^d$,
3294: is more problematic.
3295: Again we are unable to prove that the optimal path
3296: for the continuous model (i.e.\ the closure of
3297: $\bigcup_{i\in A^*} Y_i$) is connected
3298: (although we expect this to be true).
3299: In the case $d=2$ this caused a little inconvenience
3300: but we could work around it by observing that
3301: any ``jumps'' in the path consist of
3302: horizontal or vertical line segments, and hence that
3303: at least there exists a unique connected increasing path
3304: that contains the optimal path.
3305:
3306: For $d\geq 3$, however, a jump could, for example,
3307: cross a square of zero volume in $\RR^d$ but
3308: with non-zero area. Then there is no longer a unique
3309: way to extend the optimal path to a connected increasing path,
3310: and thus there is an ambiguity in the limit object.
3311: If we could prove the conjecture that the optimal path
3312: itself is connected, the convergence in distribution
3313: of the discrete optimal paths $P^{(n)*}$ would follow as before.
3314:
3315: In the case $\alpha=0$, we can in fact prove the connectedness
3316: of the optimal path for the continuous model
3317: (defined as in Section \ref{greedysection} using
3318: the ``greedy algorithm'').
3319: It is not clear how to extend the multifractal
3320: analysis of Section \ref{greedysubsection}.
3321: However, by analysing a branching
3322: random walk associated with the algorithm which constructs
3323: the greedy path, one can obtain
3324: that the function from, say, $x_1\in[0,1]$
3325: to $(x_2,\dots, x_d)\in[0,1]^{d-1}$ which describes the path
3326: is almost surely everywhere continuous and strictly increasing
3327: (although a.s.\ it also has derivative 0 almost everywhere).
3328: Thus one can show that $P^{(n)*}\to P^*$
3329: in distribution (under the Hausdorff metric on subsets of $[0,1]^d$)
3330: for all $d$ in the case $\alpha=0$.
3331:
3332: \section*{Acknowledgments}
3333: We are grateful for the support of the Isaac Newton Institute in Cambridge;
3334: this work began during the programme \textit{Interaction
3335: and Growth in Complex Stochastic Systems}.
3336:
3337: %\bibliographystyle{usual2}
3338: %\bibliography{jbm}
3339:
3340: \begin{thebibliography}{10}
3341:
3342: \bibitem{AldDiapatience}
3343: \textsc{Aldous, D. and Diaconis, P.}, (1999) Longest increasing subsequences:
3344: from patience sorting to the {B}aik-{D}eift-{J}ohansson theorem.
3345: \newblock \emph{Bull. Amer. Math. Soc. (N.S.)} \textbf{36}, 413--432.
3346:
3347: \bibitem{ap}
3348: \textsc{Arbeiter, M. and Patzschke, N.}, (1996) Random self-similar
3349: multifractals.
3350: \newblock \emph{Math. Nachr.} \textbf{181}, 5--42.
3351:
3352: \bibitem{BBM}
3353: \textsc{Baccelli, F., Borovkov, A. and Mairesse, J.}, (2000) Asymptotic
3354: results on infinite tandem queueing networks.
3355: \newblock \emph{Probab.\ Theory Related Fields} \textbf{118}, 365--405.
3356:
3357: \bibitem{BDMMZ}
3358: \textsc{Baik, J., Deift, P., McLaughlin, K. T.-R., Miller, P. and Zhou, X.},
3359: (2001) Optimal tail estimates for directed last passage site percolation with
3360: geometric random variables.
3361: \newblock \emph{Adv. Theor. Math. Phys.} \textbf{5}, 1207--1250.
3362:
3363: \bibitem{BaiSui}
3364: \textsc{Baik, J. and Suidan, T.}, (2005) A {G}{U}{E} central limit theorem and
3365: universality of directed first and last passage site percolation.
3366: \newblock \emph{Int. Math. Res. Not.} \textbf{2005:6}, 325--337.
3367:
3368: \bibitem{GUEs}
3369: \textsc{Baryshnikov, Y.}, (2001) GUEs and queues.
3370: \newblock \emph{Probab.\ Theory Related Fields} \textbf{119}, 256--274.
3371:
3372: \bibitem{Bertoinbook}
3373: \textsc{Bertoin, J.}, (1996) \emph{L\'evy processes}, vol. 121 of
3374: \emph{Cambridge Tracts in Mathematics}.
3375: \newblock Cambridge University Press, Cambridge.
3376:
3377: \bibitem{bhj}
3378: \textsc{Biggins, J.~D., Hambly, B.~M. and Jones, O.~D.}, (2006) Multifractal
3379: spectra for random self-similar measures via branching processes.
3380: \newblock In preparation.
3381:
3382: \bibitem{BodineauMartin}
3383: \textsc{Bodineau, T. and Martin, J.~B.}, (2005) A universality property for
3384: last-passage percolation paths close to the axis.
3385: \newblock \emph{Electron. Comm. Probab.} \textbf{10}, 105--112 (electronic).
3386:
3387: \bibitem{Davidbook}
3388: \textsc{David, H.~A.}, (1981) \emph{Order statistics}.
3389: \newblock John Wiley \& Sons Inc., New York, 2nd ed.
3390: \newblock Wiley Series in Probability and Mathematical Statistics.
3391:
3392: \bibitem{fal}
3393: \textsc{Falconer, K.~J.}, (1994) The multifractal spectrum of statistically
3394: self-similar measures.
3395: \newblock \emph{J. Theoret. Probab.} \textbf{7}, 681--702.
3396:
3397: \bibitem{GlyWhi}
3398: \textsc{Glynn, P.~W. and Whitt, W.}, (1991) Departures from many queues in
3399: series.
3400: \newblock \emph{Ann.\ Appl.\ Probab.} \textbf{1}, 546--572.
3401:
3402: \bibitem{GraTraWid}
3403: \textsc{Gravner, J., Tracy, C.~A. and Widom, H.}, (2001) Limit theorems for
3404: height fluctuations in a class of discrete space and time growth models.
3405: \newblock \emph{J.\ Statist.\ Phys.} \textbf{102}, 1085--1132.
3406:
3407: \bibitem{hmo}
3408: \textsc{Hambly, B.~M., Martin, J.~B. and O'Connell, N.}, (2002) Concentration
3409: results for a {B}rownian directed percolation problem.
3410: \newblock \emph{Stochastic Process.\ Appl.} \textbf{102}, 207--220.
3411:
3412: \bibitem{JLRbook}
3413: \textsc{Janson, S., {\L}uczak, T. and Rucinski, A.}, (2000) \emph{Random
3414: graphs}.
3415: \newblock Wiley-Interscience Series in Discrete Mathematics and Optimization.
3416: Wiley-Interscience, New York.
3417:
3418: \bibitem{Johshape}
3419: \textsc{Johansson, K.}, (2000) Shape fluctuations and random matrices.
3420: \newblock \emph{Comm. Math. Phys.} \textbf{209}, 437--476.
3421:
3422: \bibitem{Johanssontransversal}
3423: \textsc{Johansson, K.}, (2000) Transversal fluctuations for increasing
3424: subsequences on the plane.
3425: \newblock \emph{Probab. Theory Related Fields} \textbf{116}, 445--456.
3426:
3427: \bibitem{JohanssonAiry}
3428: \textsc{Johansson, K.}, (2003) Discrete polynuclear growth and determinantal
3429: processes.
3430: \newblock \emph{Comm. Math. Phys.} \textbf{242}, 277--329.
3431:
3432: \bibitem{KrugSpohn}
3433: \textsc{Krug, J. and Spohn, H.}, (1992) Kinetic roughening of growing
3434: surfaces.
3435: \newblock In C.~Godr{\`e}che, ed., \emph{Solids far from equilibrium},
3436: Collection Al\'ea-Saclay: Monographs and Texts in Statistical Physics, 1,
3437: pages 479--582. Cambridge University Press, Cambridge.
3438:
3439: \bibitem{blocking}
3440: \textsc{Martin, J.~B.}, (2002) Large tandem queueing networks with blocking.
3441: \newblock \emph{Queueing Syst. Theory Appl.} \textbf{41}, 45--72.
3442:
3443: \bibitem{animals}
3444: \textsc{Martin, J.~B.}, (2002) Linear growth for greedy lattice animals.
3445: \newblock \emph{Stochastic Process.\ Appl.} \textbf{98}, 43--66.
3446:
3447: \bibitem{JBMshape}
3448: \textsc{Martin, J.~B.}, (2004) Limiting shape for directed percolation models.
3449: \newblock \emph{Ann. Probab.} \textbf{32}, 2908--2937.
3450:
3451: \bibitem{Neilsurvey}
3452: \textsc{O'Connell, N.}, (2003) Random matrices, non-colliding particle systems
3453: and queues.
3454: \newblock In \emph{S\'{e}minaire de Probabilit\'{e}s XXXVI}, no. 1801 in
3455: Lecture Notes in Mathematics, pages 165--182. Springer-Verlag.
3456:
3457: \bibitem{OcoYor}
3458: \textsc{O'Connell, N. and Yor, M.}, (2001) Brownian analogues of {B}urke's
3459: theorem.
3460: \newblock \emph{Stochastic Process. Appl.} \textbf{96}, 285--304.
3461:
3462: \bibitem{PraSpo}
3463: \textsc{Pr{\"a}hofer, M. and Spohn, H.}, (2002) Scale invariance of the {PNG}
3464: droplet and the {A}iry process.
3465: \newblock \emph{J. Statist. Phys.} \textbf{108}, 1071--1106.
3466:
3467: \bibitem{Resnick86}
3468: \textsc{Resnick, S.~I.}, (1986) Point processes, regular variation and weak
3469: convergence.
3470: \newblock \emph{Adv. in Appl. Probab.} \textbf{18}, 66--138.
3471:
3472: \bibitem{Rost}
3473: \textsc{Rost, H.}, (1981) Nonequilibrium behaviour of a many particle process:
3474: density profile and local equilibria.
3475: \newblock \emph{Z. Wahrsch. Verw. Gebiete} \textbf{58}, 41--53.
3476:
3477: \bibitem{Soshnikovheavy}
3478: \textsc{Soshnikov, A.}, (2004) Poisson statistics for the largest eigenvalues
3479: of {W}igner random matrices with heavy tails.
3480: \newblock \emph{Electron. Comm. Probab.} \textbf{9}, 82--91 (electronic).
3481:
3482: \end{thebibliography}
3483:
3484:
3485: \bigskip
3486:
3487: \parbox{0.52\linewidth}
3488: {
3489: \textsc{Mathematical Institute,\\
3490: University of Oxford,\\
3491: 24-29 St Giles,\\
3492: Oxford OX1 3LB,\\
3493: UK}\\
3494: \texttt{hambly@maths.ox.ac.uk}\\
3495: \texttt{http://www.maths.ox.ac.uk/$\tilde{\,\,\,\,}$hambly}}
3496: \hfill
3497: \parbox{0.47\linewidth}
3498: {\textsc{Department of Statistics,\\
3499: University of Oxford,\\
3500: 1 South Parks Road,\\
3501: Oxford OX1 3TG\\
3502: UK}\\
3503: \texttt{martin@stats.ox.ac.uk}\\
3504: \texttt{http://www.stats.ox.ac.uk/$\tilde{\,\,\,\,}$martin}}
3505:
3506:
3507: \end{document}
3508:
3509: