math0603306/third
1: %06.03.13
2: %This version is considered final on March 13. It goes to arxiv.
3: \documentclass{article}
4: %\documentclass[draft]{article}
5: \usepackage[leqno,intlimits]{amsmath}
6: \usepackage{amssymb}
7: \usepackage{amsthm}
8: %\usepackage{showkeys}
9: \usepackage{hyperref}
10: \allowdisplaybreaks[1]
11: 
12: \newtheorem{tm}{Theorem}[section]
13: \newtheorem{lm}[tm]{Lemma}
14: \newtheorem{cor}[tm]{Corollary}
15: 
16: 
17: \newcommand{\va}{a}  %%% temporary parameter
18: 
19: \newcommand*{\hop}{\bigskip\noindent}
20: \newcommand{\nn}{\nonumber}
21: \newcommand*{\Rb}{\mathbb R}
22: \newcommand*{\Zb}{\mathbb Z}
23: \newcommand*{\un}{\underline}
24: \newcommand*{\ba}{\begin{aligned}}
25: \newcommand*{\ea}{\end{aligned}}
26: \newcommand*{\be}{\begin{equation}}
27: \newcommand*{\ee}{\end{equation}}
28: \newcommand*{\vr}{\varrho}
29: \newcommand*{\vp}{\varphi}
30: \newcommand*{\ve}{\varepsilon}
31: \newcommand*{\om}{\omega}
32: \newcommand*{\al}{\alpha}
33: \newcommand*{\la}{\lambda}
34: \newcommand*{\pt}{\partial}
35: \newcommand*{\di}{\,\text{\rm d}}
36: \newcommand*{\les}{\lesssim}
37: \newcommand*{\Ac}{\mathcal A}
38: \newcommand*{\Wc}{\mathcal W}
39: \newcommand*{\Nc}{\mathcal N}
40: \newcommand*{\Ec}{\mathcal E}
41: \newcommand*{\Sc}{\mathcal S}
42: \newcommand*{\wt}{\widetilde}
43: \newcommand*{\wh}{\widehat}
44: \newcommand*{\Ev}{\mathbf E}
45: \newcommand*{\Pv}{\mathbf P}
46: \newcommand*{\Vv}{{\text{\bf Var}}}
47: \newcommand*{\Cov}{\text{\bf Cov}}
48: \newcommand*{\e}[1]{\text{\rm e}^{#1}}
49: \newcommand*{\fl}[1]{\left\lfloor{#1}\right\rfloor}
50: \newcommand*{\tfl}[1]{\lfloor{#1}\rfloor}
51: \newcommand*{\si}{\sigma}
52: \DeclareMathOperator\Expd{Exp}
53: \numberwithin{equation}{section}
54: \begin{document}
55: \title{Cube root fluctuations for the corner growth model
56: associated to the
57: exclusion process}
58: \author{M.\ Bal\'azs\thanks{University of Wisconsin-Madison\newline
59: M. Bal\'azs was partially supported by the Hungarian Scientific Research Fund
60: (OTKA) grant TS49835 and by National Science Foundation Grant DMS-0503650.\newline
61: T.\ Sepp\"al\"ainen was partially supported by National Science Foundation
62: grant DMS-0402231.}, E. Cator\thanks{Delft University of Technology}
63:  and  T.\ Sepp\"al\"ainen$^*$}
64: \maketitle
65: \begin{abstract}
66: We study the last-passage growth model on the planar integer lattice
67:  with exponential weights.  With boundary conditions
68:  that represent the equilibrium exclusion process as seen
69: from a particle right after its jump
70: we prove that the variance of the last-passage time
71: in a characteristic direction
72: is of order $t^{2/3}$.  With more general boundary
73: conditions that include the rarefaction fan case
74: we show that the last-passage time fluctuations are
75: still  of order $t^{1/3}$, and also that
76: the transversal fluctuations of the maximal path
77: have order $t^{2/3}$.
78: We   adapt and then build on   a recent study
79: of Hammersley's process by Cator and Groeneboom, and also
80: utilize   the competition interface introduced by Ferrari,
81: Martin and Pimentel.
82:   The arguments are entirely probabilistic, and no
83: use is made of the combinatorics of Young tableaux or methods of
84: asymptotic analysis.
85: \end{abstract}
86: 
87: \noindent
88: {\bf Keywords:} Last-passage, simple exclusion, cube root asymptotics, competition interface, Burke's theorem, rarefaction fan
89: 
90: \hop
91: {\bf MSC:} 60K35, 82C43
92: 
93: \section{Introduction}
94: 
95: We construct a version of the corner growth model that corresponds
96: to an equilibrium exclusion process as seen by a typical particle
97: right after its jump,
98: and show that along a characteristic direction the variance of the
99: last-passage time is of order $t^{2/3}$.  This last-passage time
100: is the maximal sum of  exponential weights along up-right paths
101: in the first quadrant of the integer plane.
102: The interior weights have rate 1, while the boundary weights
103: on the axes have rates
104: $1-\vr$ and $\vr$ where $0<\vr<1$ is the particle density of the
105: exclusion process.
106: By comparison to this equilibrium setting, we also show
107: fluctuation results with
108:  similar scaling in the case of the rarefaction fan.
109: 
110: The proof is based on a recent work of Cator and Groeneboom
111: \cite{cuberoot} where corresponding results are proved for
112: the planar-increasing-path version of
113: Ham\-mers\-ley's process.  A key part of that proof is an identity
114: that relates the variance of the last-passage time to the
115: point where the maximal path exits the axes. This exit point itself
116: is related to a second-class particle via a time reversal.
117: The idea that the current and  the second-class particle
118: should be connected goes  back to a paper of Ferrari and Fontes
119: \cite{se} on the diffusive fluctuations
120: of the current away from the characteristic.  However, despite this
121: surprising congruence of ideas, article  \cite{cuberoot} and our
122: work have  no technical relation to the
123: Ferrari-Fontes work.
124: 
125: The first task of the present paper is to find the
126:  connection between the variance of the last-passage time and
127: the  exit point, in the equilibrium corner growth model.  The relation
128: turns out not as straightforward as for Hammersley's process,
129: for we also need to include the amount of weight collected on
130: the axes.  However, once this difference is understood, the arguments
131: proceed quite similarly to those in  \cite{cuberoot}.
132: 
133: The notion of competition interface recently introduced
134: by Ferrari, Martin and Pimentel \cite{fermarpim, compint} now appears as
135: the representative of a second-class particle, and as the time
136: reversal of the maximal path.  As a by-product of the proof we
137: establish that the transversal fluctuations of the competition
138: interface are of the order
139: $t^{2/3}$ in the equilibrium setting.
140: 
141: In the last section we take full advantage of our
142: probabilistic approach, and show that for initial conditions obtained
143: by decreasing the equilibrium  weights on the axes
144: in an {\sl arbitrary} way,
145:  the fluctuations
146: of the last-passage time are still of order $t^{1/3}$.
147: This includes the situation known as the {\sl rarefaction fan}.
148: We are also able
149: to show that in this case the transversal fluctuations of the longest
150: path are of order $t^{2/3}$. In this more general setting
151: there is no direct connection between a maximal path and a competition
152: interface (or trajectory of a second class particle).
153: 
154: 
155: Our results for the competition interface, and
156: our fluctuation results  under the
157: more general boundary conditions are new.
158: The variance bound for the equilibrium last-passage time
159: is also strictly speaking new. However,
160:  the corresponding
161: distributional limit has been obtained by Ferrari and Spohn
162: \cite{ferspohn} with a  proof based on the RSK machinery.
163: But they lack a suitable tightness property that would give them
164:  also control of the variance. [Note that
165: Ferrari and Spohn start by describing a different set of
166: equilibrium boundary conditions than the ones we consider, but later
167: in their paper they cover also the kind we  define in
168: \eqref{eq:bondis} below.]
169: The methods of our paper can  also be applied to
170: geometrically distributed weights, with the same outcomes.
171: 
172: %The distributional and variance
173: %limits of the last-passage time in the equilibrium corner growth model
174: % have already % been derived by Ferrari and Spohn \cite{ferspohn}.
175: %Although they start with slightly different initial conditions,
176: %their work also covers our boundary conditions for the
177: %equilibrium setting.
178: 
179: In addition to the results themselves,
180: our main  motivation  is to investigate
181:  new methods to attack the last-passage model, methods
182: that do not rely on the RSK correspondence of Young tableaux.
183: The reason for such a pursuit is that the precise counting
184: techniques of Young tableaux appear to work only for geometrically
185: distributed weights, from which one can then take a limit
186: to  obtain the case of  exponential weights. New techniques are needed
187: to go beyond the geometric and exponential cases,
188: although we are not yet in a position to undertake such an advance.
189: 
190: %%In the present paper we work directly with
191: %%exponential weights and obtain results for specific boundary
192: %conditions which are relevant to the equilibrium simple exclusion
193: %process, and then for more general rarefaction boundary conditions
194: %as well.
195: 
196: For the  class of totally asymmetric stochastic interacting
197: systems for which the last-passage approach  works, this point of view
198:  has been extremely valuable.  In addition to the papers already
199: mentioned above, we list
200:  Sepp\"al\"ainen \cite{timogrowth, hkl}, Johansson \cite{1/3}, and
201:  Pr\"ahofer and Spohn  \cite{spohn}.
202: %%% Ferrari, Martin and Pimentel \cite{fermarpim}, \cite{compint}.
203: %Among these papers,
204: %\cite{ferspohn} and \cite{spohn} have applied the last-passage approach
205: % in an equilibrium situation.
206: 
207: {\bf Organization of the paper.}  The main results are discussed
208: in Section \ref{sc:recults}. Section \ref{sc:part}
209: describes the relationship of the last-passage model to
210: particle and deposition models, and can be skipped without loss
211: of continuity.
212: The remainder of the paper is for the proofs.  Section \ref{sec:prel}
213: covers some preliminary matters. This includes a strong form
214: of Burke's theorem for the last-passage times (Lemma \ref{lm:NE}).
215: Upper and lower bounds for the equilibrium results are
216: covered in Sections \ref{sc:ub} and \ref{sc:lb}.
217: Lastly, fluctuations under more
218: general boundary conditions are studied in Section \ref{sc:gen-bd}.
219: 
220: {\bf Notation.}  $\Zb_+=\{0,1,2,\dotsc\}$ denotes the set
221: of nonnegative integers.  The integer part of a real number
222: is $\tfl{x}=\max\{ n\in\Zb: n\leq x\}$.
223: $C$ denotes constants whose precise value is immaterial and that do not depend
224: on the parameter (typically $t$) that grows.
225: $X\sim \Expd(\vr)$ means that $X$ has the exponential distribution with rate
226: $\vr$, in other words has density  $f(x)=\vr e^{-\vr x}$ on $\Rb_+$.
227:   For clarity,
228: subscripts can be replaced
229: by arguments in parentheses, as for example in $G_{ij}=G(i,j)$.
230: 
231: \section{Results}
232: \label{sc:recults}
233: We start by describing the corner growth model with boundaries
234: that correspond to a special view of the equilibrium.
235: Section \ref{sc:part} and Lemma \ref{lm:NE} justify the term equilibrium in this context. Our results for more general
236:  boundary conditions are in Section \ref{sc:rareres}.
237: 
238: \subsection{Equilibrium results}
239: 
240:   We are given an array
241: $\{\om_{ij}\}_{i,j\in\Zb_+}$ of nonnegative real numbers.
242: We will always have $\om_{00}=0$.
243: The values $\om_{ij}$ with either $i=0$ or $j=0$ are
244: the boundary values, while $\{\om_{ij}\}_{i,j\geq 1}$ are
245: the interior values.
246: 
247:  Figure \ref{fig:iniri} depicts this
248: initial set-up on the first quadrant $\Zb_+^2$ of the integer plane.
249: A $\star$ marks $(0,\,0)$,
250: $\triangledown$'s mark positions $(i,\,0)$, $i\geq1$, $\vartriangle$'s
251: positions  $(0,\,j)$, $j\geq1$, and interior points $(i,\,j)$, $i,j\geq1$
252: are marked  with $\circ$'s.  The coordinates of a few
253: points around $(5,\,2)$ have been labeled.
254: 
255: For a point $(i,\,j)\in\Zb_+^2$, let $\Pi_{ij}$
256: be the set of directed paths
257: \be
258: \pi=\{(0,\,0)=(p_0,\,q_0)\to(p_1,\,q_1)\to\dots\to(p_{i+j},\,q_{i+j})=(i,\,j)\}
259: \label{eq:pi-1} \ee
260: with up-right steps
261: \be
262: (p_{l+1},\,q_{l+1})-(p_l,\,q_l)= (1,0) \ \text{ or } \ (0,1)
263: \label{eq:allow}
264: \ee
265: along the coordinate directions.
266: Define the \emph{last passage time} of the point $(i,\,j)$ as
267: \[
268: G_{ij}=\max_{\pi\in\Pi_{ij}}\sum_{(p,q)\in\pi}\om_{pq}.
269: \]
270:  $G$ satisfies the recurrence
271: \be
272: G_{ij}=(G_{\{i-1\}j}\lor G_{i\{j-1\}})+\om_{ij}\qquad(i,\,j\geq0)\label{eq:grec}
273: \ee
274: (with formally assuming $G_{\{-1\}j}=G_{i\{-1\}}=0$).
275: A common interpretation is that this models a  growing cluster
276: on the
277: first quadrant that starts from a seed at the origin
278: (bounded by the thickset line in Figure \ref{fig:iniri}).
279: The value  $\om_{ij}$
280: is the time it takes to occupy point $(i,j)$  after its neighbors
281: to the left and below have become occupied, with the interpretation
282: that a boundary point needs only one occupied neighbor.
283: Then $G_{ij}$ is the time when $(i,j)$ becomes occupied, or
284:  joins the growing cluster.
285: The occupied region at time  $t\geq0$ is the set
286: \be
287: \Ac(t)=\{(i,j)\in\Zb_+^2\,:\, G_{ij}\leq t\}.
288: \label{eq:def-Ac}
289: \ee
290: 
291: Figure \ref{fig:laterri} shows a possible later situation.
292: Occupied points are denoted by solidly colored symbols, the
293: occupied cluster is bounded by the thickset line, and the
294: arrows mark an admissible path $\pi$ from $(0,0)$ to $(5,2)$.
295: %%If $G_{4,2}>G_{5,1}$ then this path is actually the maximal path to $(5,2)$.
296: If $G_{5,2}$ is the smallest among
297: $G_{0,5}$, $G_{1,4}$, $G_{5,2}$ and $G_{6,0}$, then $(5,2)$
298: is the next point added to the cluster, as suggested by the
299: dashed lines around the $(5,2)$ square.
300: 
301: To create a model of random evolution, we pick  a real number
302: $0<\vr<1$ and  take  the variables $\{\om_{ij}\}$  mutually independent
303: with the following marginal distributions:
304: \be
305: \ba
306: \om_{00}&=0,\qquad&&\text{where the }\star\text{ is},\\
307: \om_{i0}&\sim\Expd(1-\vr),\ i\geq1,\qquad&&\text{where the }\triangledown\text{'s are},\\
308: \om_{0j}&\sim\Expd(\vr),\ j\geq1,\qquad&&\text{where the }\vartriangle\text{'s are},\\
309: \om_{ij}&\sim\Expd(1),\ i,j\geq1,\qquad&&\text{where the }\circ\text{'s are}.
310: \ea\label{eq:bondis}
311: \ee
312: 
313: %%%%%%%%%%%%1 start
314: \begin{figure}[ht]
315: \begin{center}
316: \begin{picture}(220,140)(5,20)
317: \put(40,30){\vector(1,0){150}}
318: \put(50,20){\vector(0,1){130}}
319: \multiput(50,40)(0,20){6}{\line(1,0){140}}
320: \multiput(60,30)(20,0){7}{\line(0,1){120}}
321: \put(47,27){\large$\star$}
322: \multiput(66,26)(20,0){6}{\large$\triangledown$}
323: \multiput(45.6,47)(0,20){5}{\large$\vartriangle$}
324: \multiput(67,47)(0,20){5}{
325: \multiput(0,0)(20,0){6}{\large$\circ$}}
326: \put(32,23){\tiny$(0,0)$}
327: \put(185,23){\tiny$i$}
328: \put(43,145){\tiny$j$}
329: \put(69,21){\tiny$1$}
330: \put(89,21){\tiny$2$}
331: \put(109,21){\tiny$3$}
332: \put(129,21){\tiny$4$}
333: \put(149,21){\tiny$5$}
334: \put(169,21){\tiny$6$}
335: \put(42,47){\tiny$1$}
336: \put(42,67){\tiny$2$}
337: \put(42,87){\tiny$3$}
338: \put(42,107){\tiny$4$}
339: \put(42,127){\tiny$5$}
340: \put(142,63){\tiny$(5,2)$}
341: \put(142,43){\tiny$(5,1)$}
342: \put(142,83){\tiny$(5,3)$}
343: \put(122,63){\tiny$(4,2)$}
344: \put(162,63){\tiny$(6,2)$}
345: {\linethickness{2pt}
346: \put(50,40){\line(1,0){10}}
347: \put(60,30){\line(0,1){10}}
348: }
349: \end{picture}
350: \end{center}
351: \caption{The initial situation}\label{fig:iniri}
352: \end{figure}%
353: 
354: 
355: 
356: \begin{figure}[ht]
357: \begin{center}
358: \begin{picture}(220,140)(5,20)
359: \put(40,30){\vector(1,0){150}}
360: \put(50,20){\vector(0,1){130}}
361: \multiput(50,40)(0,20){6}{\line(1,0){140}}
362: \multiput(60,30)(20,0){7}{\line(0,1){120}}
363: \put(47,27){\large$\star$}
364: \multiput(66,26)(20,0){5}{\large$\blacktriangledown$}
365: \put(166,26){\large$\triangledown$}
366: \multiput(45.6,47)(0,20){4}{\large$\blacktriangle$}
367: \put(45.6,127){\large$\vartriangle$}
368: \multiput(67,47)(0,20){3}{
369: \multiput(0,0)(20,0){4}{\large$\bullet$}}
370: \put(147,47){\large$\bullet$}
371: \multiput(67,107)(0,20){2}{
372: \multiput(0,0)(20,0){6}{\large$\circ$}}
373: \multiput(147,67)(0,20){2}{\large$\circ$}
374: \multiput(167,47)(0,20){3}{\large$\circ$}
375: \put(32,23){\tiny$(0,0)$}
376: \put(185,23){\tiny$i$}
377: \put(43,145){\tiny$j$}
378: \put(69,21){\tiny$1$}
379: \put(89,21){\tiny$2$}
380: \put(109,21){\tiny$3$}
381: \put(129,21){\tiny$4$}
382: \put(149,21){\tiny$5$}
383: \put(169,21){\tiny$6$}
384: \put(42,47){\tiny$1$}
385: \put(42,67){\tiny$2$}
386: \put(42,87){\tiny$3$}
387: \put(42,107){\tiny$4$}
388: \put(42,127){\tiny$5$}
389: \put(142,63){\tiny$(5,2)$}
390: \put(142,43){\tiny$(5,1)$}
391: \put(142,83){\tiny$(5,3)$}
392: \put(122,63){\tiny$(4,2)$}
393: \put(162,63){\tiny$(6,2)$}
394: \multiput(55,33)(60,37){2}{
395: \multiput(0,0)(20,0){2}{\vector(1,0){11}}}
396: \multiput(90,35)(20,20){2}{\vector(0,1){10}}
397: \put(95,50){\vector(1,0){11}}
398: {\linethickness{2pt}
399: \put(50,120){\line(1,0){10}}
400: \put(60,100){\line(0,1){20}}
401: \put(60,100){\line(1,0){80}}
402: \put(140,60){\line(0,1){40}}
403: \put(140,60){\line(1,0){20}}
404: \put(160,30){\line(0,1){30}}
405: \multiput(141,80)(4,0){5}{\line(1,0){2}}
406: \multiput(160,61)(0,4){5}{\line(0,1){2}}
407: }
408: \end{picture}
409: \end{center}
410: \caption{A possible  later situation}\label{fig:laterri}
411: \end{figure}%
412: 
413: %%%%%%%1 end of figure
414: 
415: Ferrari, Pr\"ahofer and Spohn \cite{ferspohn}, \cite{spohn} consider the Bernoulli-equilibrium of simple exclusion, which corresponds to a slightly more complicated boundary distribution than the one described above. However, Ferrari and Spohn \cite{ferspohn} early on turn to the distribution described by \eqref{eq:bondis}, as it is more natural for last-passage. We will greatly exploit the simplicity of \eqref{eq:bondis} in Section \ref{sec:prel}. In fact, \eqref{eq:bondis} is also connected with the stationary exclusion process of particle density $\vr$. To see this point, we need to look at a particle of simple exclusion in a specific manner that we explain below in Section \ref{sec:tasep}.
416: 
417: Once the parameter $\vr$ has been picked we denote the last-passage
418: time of point $(m,n)$ by $G^\vr_{mn}$.
419: In order to see interesting behavior we follow
420: the last-passage time  along the  ray defined by
421: \be
422: (m(t),n(t))=(\tfl{(1-\vr)^2t}, \tfl{\vr^2t})
423: \label{eq:char-mn}
424: \ee
425:  as $t\to\infty$.  In Section \ref{sec:char}   we give a
426: heuristic justification for this choice. It represents the
427: characteristic speed of the macroscopic equation of the system.   Let us
428: abbreviate
429: \[
430: G^\vr(t)=G^\vr\bigl(\tfl{(1-\vr)^2t}, \tfl{\vr^2t} \bigr).
431: \]
432: Once we have proved that all horizontal and vertical increments
433: of $G$-values are distributed exponentially like the boundary increments,
434:  we see that
435: \[ \Ev(G^\vr(t))=\frac{\tfl{(1-\vr)^2t}}{1-\vr}+\frac{\tfl{\vr^2t}}{\vr}.
436: \]
437: The first result is the order of the variance.
438: 
439: \begin{tm}\label{tm:G-var-1}
440: With $0<\vr<1$
441: and  independent $\{\om_{ij}\}$ distributed as in {\rm \eqref{eq:bondis}},
442: \[
443: 0<\liminf_{t\to\infty} \frac{\Vv(G^\vr(t))}{t^{2/3}} \leq
444: \limsup_{t\to\infty} \frac{\Vv(G^\vr(t))}{t^{2/3}} <\infty.
445: \]
446: \end{tm}
447: 
448: For given $(m,n)$  there is almost
449: surely a unique path $\wh\pi$ that maximizes the
450: passage time to $(m,n)$,
451: due to the continuity of the distribution of
452: $\{\om_{ij}\}$.
453: The {\sl exit point} of $\wh\pi$ is the last boundary point
454: on the path. If $(p_l,q_l)$ is the exit point for the path
455: in  \eqref{eq:pi-1}, then either $p_0=p_1=\dotsm=p_l=0$
456: or  $q_0=q_1=\dotsm=q_l=0$, and $p_k,q_k\geq 1$ for all $k>l$.
457: To distinguish between exits via the $i$- and $j$-axis, we introduce
458: a non-zero integer-valued random variable $Z$ such that if
459: $Z>0$ then the exit point is $(p_{\lvert{Z}\rvert},q_{\lvert{Z}\rvert})=(Z,0)$,
460: while  if
461: $Z<0$ then the exit point is $(p_{\lvert{Z}\rvert},q_{\lvert{Z}\rvert})
462: =(0,-Z)$.  For the sake of convenience we abuse language and call
463: the variable $Z$ also the ``exit point.''
464:  $Z^\vr(t)$ denotes the exit point of the maximal path
465: to the point $(m(t),n(t))$ in \eqref{eq:char-mn}
466:  with
467: boundary condition parameter $\vr$.
468: Transposition $\om_{ij}\mapsto\om_{ji}$ of the array shows
469: that $Z^\vr(t)$ and $-Z^{1-\vr}(t)$ are equal in distribution.
470: Along the way to Theorem \ref{tm:G-var-1} we establish
471: that $Z^\vr(t)$ fluctuates on the scale $t^{2/3}$.
472: 
473: \begin{tm} Given $0<\vr<1$ and independent $\{\om_{ij}\}$ distributed
474:  as in {\rm \eqref{eq:bondis}}.
475: 
476: {\rm (a)} For $t_0>0$
477: there exists a finite constant  $C=C(t_0, \vr)$
478: such that, for all $a>0$ and $t\geq t_0$,
479: \[
480: \Pv\{ Z^\vr(t)\geq at^{2/3}\} \leq C a^{-3}.
481: \]
482: 
483: {\rm (b)} Given $\ve>0$, we can choose a
484:  $\delta>0$ small enough so that for all large enough  $t$
485: \[
486: \Pv\{1\leq  Z^\vr(t)\leq \delta t^{2/3}\}\leq \ve.
487: \]
488: \label{tm:Z-1}
489: \end{tm}
490: 
491: \hop
492: {\bf Competition interface.}
493: In \cite{fermarpim, compint} Ferrari, Martin and Pimentel introduced the {\sl competition
494: interface} in the last-passage picture. This is a path
495: $k\mapsto \vp_k\in\Zb_+^2$ ($k\in\Zb_+$),
496: defined as a function of $\{G_{ij}\}$:  first $\vp_0=(0,0)$,
497: and then for $k\geq 0$
498: \be
499: \vp_{k+1}= \begin{cases}
500: \vp_k+(1,0) &\text{if $G(\vp_k+(1,0))<G(\vp_k+(0,1))$,}\\
501: \vp_k+(0,1) &\text{if $G(\vp_k+(1,0))>G(\vp_k+(0,1))$.}
502: \end{cases}\label{eq:def-vphi}
503: \ee
504: In other words, $\vp$ takes up-right steps, always choosing
505: the smaller of the two possible $G$-values.
506: 
507: The term ``competition interface'' is justified by the following picture.
508: Instead of having the unit squares centered at the integer points
509: as in Figure \ref{fig:iniri},  draw the squares so that
510: their corners coincide with integer points.
511: Label the  squares  by their northeast
512: corners, so that  the square
513: $(i-1,i]\times(j-1,j]$ is labeled  the
514: $(i,j)$-square. Regard  the last-passage time $G_{ij}$
515: as the time when the $(i,j)$-square becomes occupied.  Color the square $(0,0)$
516: white. Every other square gets either a red or a blue color:
517:  squares   to
518: the left and above the path $\vp$ are colored red, and squares   to
519: the right and below  $\vp$ blue.  Then the red squares are those whose
520: maximal path $\wh\pi$ passes through $(0,1)$, while the blue squares are
521:  those whose
522: maximal path $\wh\pi$ passes through $(1,0)$.  These can be regarded as
523: two competing ``infections'' on the $(i,j)$-plane, and $\vp$ is the
524: interface between them.
525: 
526: The competition interface represents the evolution of a second-class
527: particle, and macroscopically it follows the characteristics.
528: This was one of the main points for \cite{compint}.  In the
529: present setting the competition interface
530:  is the time reversal
531: of the maximal path $\wh\pi$, as we explain more precisely
532: in Section \ref{sec:prel}  below. This
533: connection allows us to establish the order of the transversal
534: fluctuations of the competition interface in the equilibrium setting.
535: %, which we record in the next theorem.
536: To put this in precise notation, we introduce
537: \be
538: v(n)=\inf\{i\,:\,(i,n)=\vp_k\text{ for some }k\geq 0\}
539: \label{eq:def-vn}
540: \ee
541: and
542: \[
543: w(m)=\inf\{j\,:\,(m,j)=\vp_k\text{ for some }k\geq 0\}
544: \]
545: with the usual convention  $\inf\emptyset=\infty$. In other words,
546: $(v(n),n)$ is the leftmost point of the competition interface on
547: the horizontal line $j=n$, while $(m,w(m))$ is the lowest such
548: point on the vertical line $i=m$.  They are connected by the
549: implication
550: \be
551: v(n)\geq m\ \Longrightarrow \ w(m)<n
552: \label{eq:v-w-ineq}\ee
553: as can be seen from a picture.
554: Transposition $\om_{ij}\mapsto\om_{ji}$
555: of the $\om$-array interchanges $v$ and $w$.
556: 
557: Given $m$ and $n$,
558: %write temporarily  $\vp_k=(i_k,j_k)$
559: %in terms of coordinates and let
560: %\[
561: %N(t)=\inf\{ k: \,\text{$i_k=\tfl{(1-\vr)^2t}$ or $j_k=\tfl{\vr^2t}$} \}
562: %\]
563: %so that $\vp_{N(t)}$ is the point where the competition interface
564: %first hits either the east  or the north boundary of the  rectangle
565: %$\{0,\dotsc,\tfl{(1-\vr)^2}\}\times \{0,\dotsc,\tfl{\vr^2t}\}$.
566: let
567: \be
568: Z^{*\vr}=[m-v(n)]^+-[n-w(m)]^+\label{eq:zstar}
569: %\zeta^\vr(t)= \bigl( \tfl{(1-\vr)^2t}-i_{N(t)}\bigr)
570: %+ \bigl( \tfl{\vr^2t}-j_{N(t)}\bigr)
571: \ee
572: denote the signed distance from the point $(m,\,n)$
573: %northeast corner \eqref{eq:char-mn}
574: to the point where $\vp_k$ first hits either of the lines $j=n$ ($Z^{*\vr}>0$) or $i=m$ ($Z^{*\vr}<0$). Precisely one of the two terms
575:  contributes to the difference.
576: %$\vp_{N(t)}$.
577: When we let $m=m(t)$ and $n=n(t)$ according to \eqref{eq:char-mn}, we have the $t$-dependent version $Z^{*\vr}(t)$.
578: Time reversal will show that in distribution
579: $Z^{*\vr}(t)$ is equal to $Z^\vr(t)$. (The notation $Z^*$ is used in anticipation of this time reversal connection.) Consequently
580: \begin{cor}
581: Theorem \ref{tm:Z-1} is true word for word when $Z^\vr(t)$ is replaced by $Z^{*\vr}(t)$.
582: \end{cor}
583: %is result:
584: %
585: %\begin{tm} Given $0<\vr<1$ and independent $\{\om_{ij}\}$ distributed
586: % as in {\rm \eqref{eq:bondis}}.
587: %
588: %{\rm (a)} Given $t_0>0$
589: %there exists a finite constant  $C=C(t_0, \vr)$
590: %such that, for all $a>0$ and $t\geq t_0$,
591: %\[
592: %\Pv\{ \zeta^\vr(t)\geq at^{2/3}\} \leq C_4 a^{-3}.
593: %\]
594: %
595: %{\rm (b)} Given $\ve>0$, we can choose a
596: % $\delta>0$ small enough so that for all large enough  $t$
597: %\[
598: %\Pv\{\zeta^\vr(t)\leq \delta t^{2/3}\}\leq \ve.
599: %\]
600: %\label{tm:compint-1}
601: %\end{tm}
602: 
603: 
604: %It would be of great interest to study the corner growth model
605: %with different distributions for the $\om_{ij}$'s.
606: %However, except for results about behavior close to the
607: %axes \cite{},  the only rigorous results exist for
608: %exponential and geometric distributions.
609: 
610: \subsection{Results for the rarefaction fan}\label{sc:rareres}
611: 
612: We now partially generalize the previous results to
613: arbitrary boundary conditions that are bounded by
614: the equilibrium boundary conditions of \eqref{eq:bondis}.
615: Let  $\{\om_{ij}\}$ be distributed  as in \eqref{eq:bondis}.
616: Let  $\{\hat{\om}_{ij}\}$ be another array
617: defined on the same probability space
618:  such that $\hat{\om}_{00}=0$,
619:  $\hat{\om}_{ij}=\om_{ij}$ for $i,j \geq 1$, and
620: \be
621: \hat{\om}_{i0}\leq \om_{i0} \quad\text{and}\quad \hat{\om}_{0j}\leq
622: \om_{0j} \quad \forall\ i,j\geq 1.\label{eq:rareom}
623: \ee
624: In particular, $\hat{\om}_{i0}=\hat{\om}_{0j}=0$ is admissible here.
625: Section \ref{sc:rarepart} below explains how
626: these  boundary conditions can represent the so-called
627:  rarefaction fan situation of simple exclusion.
628: 
629: Let  $\hat{G}(t)$ denote the weight of the maximal path to $(m,n)$ of
630: \eqref{eq:char-mn}, using the $\{\hat{\om}_{ij}\}$ array.
631: \begin{tm}\label{tm:rare}
632: Fix $0<\al <1$. There exists a
633: constant $C=C(\al, \vr)$ such that for all $t\ge1$ and $a>0$,
634: \[
635: \Pv \{ |\hat{G}(t)-t| > at^{1/3}\} \leq Ca^{-3\al/2}.
636: \]
637: \end{tm}
638: Define also $\hat{Z}_l(t)$ as the $i$-coordinate of the right-most point on the horizontal line $j=l$ of the right-most maximal path to $(m,n)$, and $\hat{Y}_l(t)$ as the $i$-coordinate of the left-most point on the horizontal line $j=l$ of the left-most maximal path to $(m,n)$. (In this general setting we no longer necessarily have a unique maximizing path because we have not ruled out a dependence of $\{\hat{\om}_{i0},\hat{\om}_{0j}\}$ on $\{\hat{\om}_{ij}\}_{i,j \geq 1}$.)
639: \begin{tm}\label{tm:raretrans}
640: For all $0<\al<1$ there exists $C=C(\al,\,\vr)$, such that for all $a>0$, $s\le t$ with $t\ge1$ and $(k,l)=(\fl{(1-\vr)^2s},\fl{\vr^2s})$,
641: \[
642: \Pv\{\hat{Z}_l(t)\geq k+at^{2/3}\}\leq Ca^{-3\al},\quad\text{and}\quad\Pv\{\hat{Y}_l(t)\le k-at^{2/3}\}\leq Ca^{-3\al}.
643: \]
644: \end{tm}
645: 
646: \section{Particle systems and queues}\label{sc:part}
647: 
648: The proofs in our paper will only use the last-passage description of the model. However, we would like to point out several other pictures one can attach to the last-passage
649: model.  An immediate one is the
650:  totally asymmetric simple exclusion process (TASEP).
651:   The boundary conditions \eqref{eq:bondis}  of the last-passage
652: model   correspond to  TASEP  in  equilibrium,
653: as seen by a ``typical'' particle right after its jump.
654: We also briefly discuss queues, and an augmentation of the last-passage
655: picture that describes a deposition model with column growth,
656: as in \cite{fluct}.
657: 
658: %\hop
659: \subsection{The totally asymmetric simple exclusion process}
660: \label{sec:tasep}
661: This process
662: describes particles that jump unit steps  to the right
663:  on the integer lattice $\Zb$,  subject
664: to the exclusion rule that permits at most one particle per site.
665: The state of the process is a $\{0,1\}$-valued
666: sequence
667: $\wt{\un\eta}=\{\wt\eta_x\}_{x\in\Zb}$,
668: with the interpretation that  $\wt\eta_x=1$ means that site
669: $x$ is occupied by a particle, and
670:  $\wt\eta_x=0$  that $x$ is vacant.
671: The dynamics of the process are such that each $(1,0)$ pair
672: in the state
673: becomes a $(0,1)$ pair at rate 1, independently of the rest
674: of the state. In other words, each particle jumps to a vacant
675: site on its  right at rate 1, independently of other particles.
676:  The extreme points of the set of spatially
677:  translation-invariant equilibrium   distributions
678:  of this
679: process are the Bernoulli($\vr$) distributions $\nu^\vr$ indexed
680: by particle density  $0\leq\vr\leq1$.  Under $\nu^\vr$ the
681: occupation variables $\{\wt\eta_x\}$ are i.i.d.\ with
682: mean $\Ev^{\vr}(\wt\eta_x)=\vr$.
683: 
684: 
685: 
686: The Palm distribution of a particle system describes the equilibrium
687: distribution as seen from a ``typical'' particle. For a function $f$ of
688: $\wt{\un\eta}$, the Palm-expectation is
689: \[
690: \wh\Ev^\vr(f(\wt{\un\eta}))=\frac{\Ev^\vr(f(\wt{\un\eta})\cdot\wt\eta_0)}
691: {\Ev^\vr(\wt\eta_0)}
692: \]
693: in terms of the equilibrium expectation, see e.g.\ Port and Stone
694: \cite{posto}. Due to $\wt\eta_x\in\{0,\,1\}$,  for TASEP the Palm distribution is
695: the original Bernoulli($\vr$)-equilibrium conditioned on $\wt\eta_0=1$.
696: 
697: \begin{tm}[Burke]
698: Let $\wt{\un\eta}$ be a totally asymmetric simple exclusion process started from
699: the Palm distribution (i.e.\ a particle at the origin, Bernoulli measure elsewhere).
700: Then the position of the particle started at the origin is marginally a Poisson
701: process with jump rate $1-\vr$.
702: \end{tm}
703: The theorem follows from considering the inter-particle distances as M/M/1 queues.
704: Each of these distances is geometrically distributed, which is the stationary
705: distribution for the corresponding queue. Departure processes from these queues,
706: which correspond to TASEP particle jumps, are marginally Poisson due to Burke's
707: Theorem for queues, see e.g.\ Br\'emaud \cite{bremaud} for details. The Palm
708: distribution is important in this argument, as selecting a ``typical''
709: TASEP-particle assures that the inter-particle distances (or the lengths of the
710: queues) are geometrically distributed. For instance, the first particle to the
711: left of the origin in an ordinary Bernoulli equilibrium will \emph{not} see a
712: geometric distance to the next particle on its right.
713: 
714: Shortly we will explain how the boundary conditions
715: \eqref{eq:bondis} correspond to
716:  TASEP
717: started from Bernoulli($\vr$) measure, \emph{conditioned on $\wt\eta_0(0)=0$
718: and\linebreak [4]$\wt\eta_1(0)=1$}, i.e.\ a hole at the origin and a particle at site one
719: initially.  It will be convenient to give all particles and holes labels
720: that they retain as they jump (particles to the right, holes to the
721: left).
722: The particle initially at site one is labeled $P_0$, and the hole
723: initially at the origin is  labeled $H_0$. After this, all particles
724: are labeled with integers \emph{from right to left}, and all holes
725: \emph{from
726: left to right}. The position of particle $P_j$ at time $t$ is $P_j(t)$,
727: and the position of hole $H_i$  at time $t$ is $H_i(t)$.
728: Thus initially
729: \begin{align*}
730:   \dotsm < P_{3}(0)< P_{2}(0)&<P_{1}(0) < H_0(0)=0\\
731: & <1=P_0(0)
732: < H_1(0) < H_2(0)< H_3(0) <\dotsm
733: \end{align*}
734:   Since particles never
735: jump over each other, $P_{j+1}(t)<P_j(t)$ holds  at all times $t\geq 0$,
736: and by the same  token  also
737: $H_i(t)<H_{i+1}(t)$.
738: 
739: 
740: It turns out that this perturbation of the Palm distribution does
741: not entirely spoil Burke's Theorem.
742: \begin{cor}\label{cr:burke}
743: Marginally, $P_0(t)-1$ and $-H_0(t)$
744: are two independent Poisson processes with
745: respective jump rates $1-\vr$ and $\vr$.
746: \end{cor}
747: \begin{proof}
748: The evolution of $P_0(t)$ depends only on the initial
749: configuration\linebreak[4] $\{\wt\eta_x(0)\}_{x>1}$ and the Poisson
750: clocks governing the jumps over the edges $\{x\to x+1\}_{x\geq1}$. The
751: evolution of $H_0(t)$ depends only on the initial configuration $\{\wt\eta_x(0)\}_{x<0}$
752: and the Poisson clocks governing the jumps over the edges $\{x\to x+1\}_{x<0}$.
753: Hence $P_0(t)$ and $H_0(t)$ are independent. Moreover, $\{\wt\eta_x(0)\}_{x>1,\ x<0}$
754: is Bernoulli($\vr$) distributed, just like in the Palm distribution. Hence
755: Burke's Theorem applies to $P_0(t)$. As for $H_0(t)$, notice that $\un1-\un{\wt\eta}(t)$,
756: with $\un1_{x}\equiv1$, is a TASEP with holes and particles interchanged and
757: particles jumping to the left. Hence Burke's Theorem applies to $-H_0(t)$.
758: \end{proof}%
759: 
760: Now we can state the precise connection with the last-passage model.
761: For $i,j\geq 0$ let $T_{ij}$ denote the time when particle $P_j$ and hole $H_i$ exchange
762: places, with $T_{00}=0$.  Then
763: \[
764: \text{the processes $\{G_{ij}\}_{i,j\geq0}$ and $\{T_{ij}\}_{i,j\geq0}$ are equal in
765: distribution.  }
766: \]
767: For the marginal distributions on the
768: $i$- and $j$-axes  we see the truth
769: of the statement from Corollary \ref{cr:burke}.  More generally,
770: we can compare  the growing cluster
771: \[
772: \mathcal C(t)=\{(i,j)\in\Zb_+^2: T_{ij}\leq t\}
773: \]
774: with $\Ac(t)$   defined by \eqref{eq:def-Ac},
775: and observe that they are  countable state Markov chains
776: with the same  initial state and  identical bounded jump rates.
777: 
778: Since each particle jump corresponds to exchanging places
779: with a particular hole, one can deduce that at time $T_{ij}$,
780: \be
781: P_j(T_{ij})=i-j+1 \quad\text{ and }\quad H_i(T_{ij})=i-j.\label{eq:places}
782: \ee
783: 
784: %\subsection{M/M/1 queues and a zero range process}
785: 
786: %***** Anything worth putting here? Maybe the connection with the theorem in Walrand's book about jumping customer seeing equilibrium? ***** <---- Coming later in the Markovian business.
787: 
788: %Returning to
789: By
790: the queuing interpretation of the
791: TASEP, we represent particles as ser\-vers, and the holes between $P_j$ and $P_{j-1}$
792: as customers in the queue of server $j$. Then \emph{the occupation of the
793: last-passage point $(i,\,j)$ is the same event as the
794: completion of the service of customer $i$
795: by server $j$}. This infinite system of queues is equivalent to a constant rate totally asymmetric zero range process.
796: 
797: \subsection{The rarefaction fan}\label{sc:rarepart}
798: 
799: The classical rarefaction fan initial condition for TASEP
800: is constructed with two densities $\la_\ell>\la_r$.
801: Initially  particles to the left of the origin
802: obey Bernoulli $\la_\ell$ distributions, and particles to the
803: right of the origin follow Bernoulli $\la_r$ distributions.
804: Of interest here is the behavior of a second-class particle
805: or the competition interface, and we refer the reader to
806: articles \cite{serf, compint, fermarpim, mountguiol, seck}
807: 
808: Following the development of the previous section,
809:  condition this initial
810:  measure on having a hole $H_0$ at $0$, and a particle $P_0$
811: at $1$. Then as  observed earlier, $H_0$
812: jumps to the left according to a Poisson$(\la_\ell)$ process, while
813: $P_0$ jumps to the right according to a Poisson$(1-\la_r)$ process.
814: To represent this situation in the last-passage picture,  choose
815: boundary weights $\{\hat{\om}_{i0}\}$ i.i.d.~Exp($1-\la_r$), and
816:  $\{\hat{\om}_{0j}\}$ i.i.d.~Exp($\la_\ell$),  corresponding to the waiting
817: times of $H_0$ and $P_0$. Suppose  $\la_\ell>\vr>\la_r$ and
818: $\om$ is the $\vr$-equilibrium
819: boundary condition defined by \eqref{eq:bondis}.
820: Then we have the
821: stochastic domination  $\om_{i0}\ge \hat{\om}_{i0}$ and
822:  $\om_{0j}\ge \hat\om_{0j}$, and we can realize these inequalities
823: by coupling the boundary weights.   The proofs
824: of Section \ref{sc:gen-bd} show that in fact one need not
825: insist on exponential boundary weights $\{\hat{\om}_{i0}, \hat{\om}_{0j}\}$,
826: but instead only inequality \eqref{eq:rareom} is required for
827: the fluctuations.
828: 
829: \subsection{A deposition model}
830: 
831: In this section we describe a deposition model that gives a direct graphical connection between
832: %links with
833: the
834: TASEP and the last-passage percolation.  This point of view
835: is not needed for the later proofs, hence we only give a
836: brief explanation.
837: %% outlined here can be made precise
838: %% to make the connection precise via this graphical construction.
839: 
840: \begin{figure}[ht]
841: \begin{center}
842: \begin{picture}(320,180)(-20,-20)
843: \put(-10,0){\vector(1,0){300}}
844: \put(0,-10){\vector(0,1){160}}
845: \put(140,30){\vector(1,0){150}}
846: \put(157,23){\vector(-1,1){127}}
847: \put(140,40){\line(1,0){150}}
848: \put(120,60){\line(1,0){170}}
849: \put(100,80){\line(1,0){190}}
850: \put(80,100){\line(1,0){210}}
851: \put(60,120){\line(1,0){230}}
852: \put(40,140){\line(1,0){250}}
853: \put(40,140){\line(0,1){10}}
854: \put(60,120){\line(0,1){30}}
855: \put(80,100){\line(0,1){50}}
856: \put(100,80){\line(0,1){70}}
857: \put(120,60){\line(0,1){90}}
858: \put(140,40){\line(0,1){110}}
859: \multiput(160,30)(20,0){7}{\line(0,1){120}}
860: \put(147,27){\large$\star$}
861: \multiput(166,26)(20,0){6}{\large$\triangledown$}
862: \multiput(126,47)(-20,20){5}{\large$\vartriangle$}
863: \multiput(147,47)(20,0){7}{\large$\circ$}
864: \multiput(127,67)(20,0){8}{\large$\circ$}
865: \multiput(107,87)(20,0){9}{\large$\circ$}
866: \multiput(87,107)(20,0){10}{\large$\circ$}
867: \multiput(67,127)(20,0){11}{\large$\circ$}
868: \put(139,23){\tiny$(0,0)$}
869: \put(285,23){\tiny$i$}
870: \put(28,142){\tiny$j$}
871: \put(169,21){\tiny$1$}
872: \put(189,21){\tiny$2$}
873: \put(209,21){\tiny$3$}
874: \put(229,21){\tiny$4$}
875: \put(249,21){\tiny$5$}
876: \put(269,21){\tiny$6$}
877: \put(123,45){\tiny$1$}
878: \put(103,65){\tiny$2$}
879: \put(83,85){\tiny$3$}
880: \put(63,105){\tiny$4$}
881: \put(43,125){\tiny$5$}
882: \put(202,63){\tiny$(5,2)$}
883: \put(222,43){\tiny$(5,1)$}
884: \put(182,83){\tiny$(5,3)$}
885: \put(182,63){\tiny$(4,2)$}
886: \put(222,63){\tiny$(6,2)$}
887: \multiput(20,-2)(20,0){14}{\line(0,1){4}}
888: \multiput(-2,20)(0,20){7}{\line(1,0){4}}
889: \put(285,-7){\tiny$x$}
890: \put(-7,147){\tiny$h$}
891: \put(15,-7){\tiny$-6$}
892: \put(35,-7){\tiny$-5$}
893: \put(55,-7){\tiny$-4$}
894: \put(75,-7){\tiny$-3$}
895: \put(95,-7){\tiny$-2$}
896: \put(115,-7){\tiny$-1$}
897: \put(138,-7){\tiny$0$}
898: \put(158,-7){\tiny$1$}
899: \put(178,-7){\tiny$2$}
900: \put(198,-7){\tiny$3$}
901: \put(218,-7){\tiny$4$}
902: \put(238,-7){\tiny$5$}
903: \put(258,-7){\tiny$6$}
904: \put(278,-7){\tiny$7$}
905: \put(-12,18){\tiny$-1$}
906: \put(-7,38){\tiny$0$}
907: \put(-7,58){\tiny$1$}
908: \put(-7,78){\tiny$2$}
909: \put(-7,98){\tiny$3$}
910: \put(-7,118){\tiny$4$}
911: \put(-7,138){\tiny$5$}
912: \put(137,5){\large$\circ$}
913: \put(157,5){\large$\bullet$}
914: \put(136,12){\tiny$H_0$}
915: \put(156,12){\tiny$P_0$}
916: {\linethickness{2pt}
917: \put(130,40){\line(1,0){30}}
918: \put(160,20){\line(0,1){20}}
919: \put(160,20){\line(1,0){10}}
920: }
921: \end{picture}
922: \end{center}
923: \caption{The initial configuration}\label{fig:ini}
924: \end{figure}
925: 
926: We start by tilting the $j$-axis and all the vertical columns of Figure \ref{fig:iniri} by 45 degrees, resulting in Figure \ref{fig:ini}.
927: This picture represents
928: the same initial situation as Figure \ref{fig:iniri}, but note
929: that now the $j$-coordinates must be read  in the direction $\nwarrow$.
930: (As before, some  squares are labeled with their $(i,j)$-coordinates.)
931: The $i-j$ tilted coordinate system is embedded in an $x-h$ orthogonal system.
932: 
933: \begin{figure}[ht]
934: \begin{center}
935: \begin{picture}(320,180)(-20,-20)
936: \put(-10,0){\vector(1,0){300}}
937: \put(0,-10){\vector(0,1){160}}
938: \put(140,30){\vector(1,0){150}}
939: \put(157,23){\vector(-1,1){127}}
940: \put(140,40){\line(1,0){150}}
941: \put(120,60){\line(1,0){170}}
942: \put(100,80){\line(1,0){190}}
943: \put(80,100){\line(1,0){210}}
944: \put(60,120){\line(1,0){230}}
945: \put(40,140){\line(1,0){250}}
946: \put(40,140){\line(0,1){10}}
947: \put(60,120){\line(0,1){30}}
948: \put(80,100){\line(0,1){50}}
949: \put(100,80){\line(0,1){70}}
950: \put(120,60){\line(0,1){90}}
951: \put(140,40){\line(0,1){110}}
952: \multiput(160,30)(20,0){7}{\line(0,1){120}}
953: \put(147,27){\large$\star$}
954: \multiput(166,26)(20,0){5}{\large$\blacktriangledown$}
955: \put(266,26){\large$\triangledown$}
956: \multiput(126,47)(-20,20){4}{\large$\blacktriangle$}
957: \put(46,127){\large$\vartriangle$}
958: \multiput(247,47)(20,0){2}{\large$\circ$}
959: \multiput(207,67)(20,0){4}{\large$\circ$}
960: \multiput(187,87)(20,0){5}{\large$\circ$}
961: \multiput(87,107)(20,0){10}{\large$\circ$}
962: \multiput(67,127)(20,0){11}{\large$\circ$}
963: \multiput(127,67)(-20,20){2}{\multiput(0,0)(20,0){4}{\large$\bullet$}}
964: \multiput(147,47)(20,0){5}{\large$\bullet$}
965: \put(139,23){\tiny$(0,0)$}
966: \put(285,23){\tiny$i$}
967: \put(28,142){\tiny$j$}
968: \put(169,21){\tiny$1$}
969: \put(189,21){\tiny$2$}
970: \put(209,21){\tiny$3$}
971: \put(229,21){\tiny$4$}
972: \put(249,21){\tiny$5$}
973: \put(269,21){\tiny$6$}
974: \put(123,45){\tiny$1$}
975: \put(103,65){\tiny$2$}
976: \put(83,85){\tiny$3$}
977: \put(63,105){\tiny$4$}
978: \put(43,125){\tiny$5$}
979: \multiput(20,-2)(20,0){14}{\line(0,1){4}}
980: \multiput(-2,20)(0,20){7}{\line(1,0){4}}
981: \put(285,-7){\tiny$x$}
982: \put(-7,147){\tiny$h$}
983: \put(15,-7){\tiny$-6$}
984: \put(35,-7){\tiny$-5$}
985: \put(55,-7){\tiny$-4$}
986: \put(75,-7){\tiny$-3$}
987: \put(95,-7){\tiny$-2$}
988: \put(115,-7){\tiny$-1$}
989: \put(138,-7){\tiny$0$}
990: \put(158,-7){\tiny$1$}
991: \put(178,-7){\tiny$2$}
992: \put(198,-7){\tiny$3$}
993: \put(218,-7){\tiny$4$}
994: \put(238,-7){\tiny$5$}
995: \put(258,-7){\tiny$6$}
996: \put(278,-7){\tiny$7$}
997: \put(-12,18){\tiny$-1$}
998: \put(-7,38){\tiny$0$}
999: \put(-7,58){\tiny$1$}
1000: \put(-7,78){\tiny$2$}
1001: \put(-7,98){\tiny$3$}
1002: \put(-7,118){\tiny$4$}
1003: \put(-7,138){\tiny$5$}
1004: \put(202,63){\tiny$(5,2)$}
1005: \put(222,43){\tiny$(5,1)$}
1006: \put(182,83){\tiny$(5,3)$}
1007: \put(182,63){\tiny$(4,2)$}
1008: \put(222,63){\tiny$(6,2)$}
1009: \multiput(57,5)(160,0){2}{\large$\circ$}
1010: \multiput(97,5)(20,0){4}{\large$\circ$}
1011: \put(77,5){\large$\bullet$}
1012: \multiput(177,5)(60,0){2}{\multiput(0,0)(20,0){2}{\large$\bullet$}}
1013: \put(203,6){$\dashrightarrow$}
1014: \put(203,6){$\dashleftarrow$}
1015: \put(56,12){\tiny$H_0$}
1016: \put(96,12){\tiny$H_1$}
1017: \put(116,12){\tiny$H_2$}
1018: \put(136,12){\tiny$H_3$}
1019: \put(156,12){\tiny$H_4$}
1020: \put(216,12){\tiny$H_5$}
1021: \put(76,12){\tiny$P_4$}
1022: \put(176,12){\tiny$P_3$}
1023: \put(196,12){\tiny$P_2$}
1024: \put(236,12){\tiny$P_1$}
1025: \put(256,12){\tiny$P_0$}
1026: \multiput(155,33)(20,37){2}{
1027: \multiput(0,0)(20,0){2}{\vector(1,0){11}}}
1028: \multiput(185,35)(0,20){2}{\vector(-1,1){10}}
1029: \put(175,50){\vector(1,0){11}}
1030: {\linethickness{2pt}
1031: \put(50,120){\line(1,0){30}}
1032: \put(80,100){\line(1,0){100}}
1033: \multiput(180,80)(60,-40){2}{\line(1,0){20}}
1034: \put(200,60){\line(1,0){40}}
1035: \put(260,20){\line(1,0){10}}
1036: \multiput(80,100)(160,-60){2}{\line(0,1){20}}
1037: \multiput(180,80)(20,-20){2}{\line(0,1){20}}
1038: \put(260,20){\line(0,1){20}}
1039: \multiput(201,80)(4,0){5}{\line(1,0){2}}
1040: \multiput(220,61)(0,4){5}{\line(0,1){2}}
1041: }
1042: \end{picture}
1043: \end{center}
1044: \caption{A possible move at a later time}\label{fig:later}
1045: \end{figure}
1046: 
1047: Figure \ref{fig:later} shows the later situation that corresponds
1048: to Figure \ref{fig:laterri}. As before, the thickset line is the boundary of the squares belonging to $\Ac(t)$ of \eqref{eq:def-Ac}. Whenever it makes sense, the height $h_x$ of a column $x$ is defined as the $h$-coordinate (i.e.\ the vertical height) of the thickset line above the edge $[x,\,x+1]$ on the $x$-axis. Define the increments $\eta_x=h_{x-1}-h_x$ and notice that, whenever defined, $\eta_x\in\{0,\,1\}$ due to the tilting we made. The last passage rules, converted for this picture, tell us that occupation of a new square happens at rate one unless it would violate $\eta_x\in\{0,\,1\}$ for some $x$. Moreover, one can read that the occupation of a square $(i,\,j)$ is the same event as the pair $(\eta_{i-j},\,\eta_{i-j+1})$ changing from $(1,\,0)$ to $(0,\,1)$. Comparing this to \eqref{eq:places} leads us to the conclusion that $\eta_x$, whenever defined, is the occupation variable of the simple exclusion process that corresponds to the last passage model. This way one can also conveniently include the particles ($\eta_x=1$) and holes ($\eta_x=0$) on the $x$-axis, as seen on the figures. Notice also that the time-increment $h_x(t)-h_x(0)$ is the cumulative particle current across the bond $[x,\,x+1]$.
1049: 
1050: \subsection{The characteristics}\label{sec:char}
1051: 
1052: One-dimensional conservative particle systems have the conservation law
1053: \[
1054: \pt_t\vr(t,\,x)+\pt_xf(\vr(t,\,x))=0
1055: \]
1056: under the Eulerian hydrodynamic scaling, where $\vr(t,\,x)$ is the expected particle number per site and $f(\vr(t,\,x))$ is the macroscopic particle flux around the rescaled position $x$ at the rescaled time $t$, see e.g.\ \cite{cl} for details. Disturbances of the solution propagate with the characteristic speed $f'(\vr)$.
1057: The macroscopic particle flux for TASEP is $f(\vr)=\vr(1-\vr)$,
1058: and consequently the characteristic speed is $f'(\vr)=1-2\vr$.
1059: %(This means that a disturbance in the density profile travels at speed $1-2\vr$.)
1060: Thus the characteristic curve started
1061: at the origin  is  $t\mapsto (1-2\vr)t$.  To identify the  point $(m,n)$
1062: in the last-passage picture that corresponds to th%
1063: %e  characteristic
1064: is
1065: curve, we reason  approximately. Namely, we look for $m$ and $n$ such that hole $H_m$ and particle $P_n$ interchange positions at around time $t$ and the characteristic position $(1-2\vr)t$. By time $t$, that particle $P_n$ has jumped over approximately $(1-\vr)t$ sites due to Burke's Theorem. Hence at time zero, $P_n$ is approximately at position $(1-2\vr)t-(1-\vr)t=-\vr t$. Since the particle density is $\vr$, the particle labels around this position are $n\approx\vr^2t$ at time zero. Similarly, holes travel at a speed $-\vr$, so hole $H_m$ starts from approximately $(1-2\vr)t+\vr t$. They have density $1-\vr$, which indicates $m\approx(1-\vr)^2t$.
1066: %at time $G_{0n}\approx n/\vr$ particle $P_n$ is at position $\approx -n$,
1067: %on average it travels at speed $1-\vr$ for a time $t-n/\vr$ to arrive
1068: %at point $(1-2\vr)t$, and so we get  the equation
1069: %\[
1070: %-n + (1-\vr)\bigl(t-\frac{n}\vr \bigr)= (1-2\vr)t.
1071: %\]
1072: %From this we  deduce $n=\vr^2t$, and then from
1073: %$m-n=(1-2\vr)t$ comes  $m=(1-\vr)^2t$.
1074: Thus we are led to consider the point
1075: $(m,n)=(\tfl{(1-\vr)^2t}, \tfl{\vr^2t})$ as done in \eqref{eq:char-mn}.
1076: 
1077: \section{Preliminaries}
1078: \label{sec:prel}
1079: 
1080: We turn to establish some basic facts and tools.
1081: First an extension
1082: of  Corollary \ref{cr:burke}
1083: to show that Burke's Theorem holds for every hole and particle
1084:  in the last-passage picture. Define
1085: \[
1086: \ba
1087: I_{ij}:&=G_{ij}-G_{\{i-1\}j}\quad\text{for }i\geq1,\ j\geq0,\quad\text{ and}\\
1088: J_{ij}:&=G_{ij}-G_{i\{j-1\}}\quad\text{for }i\geq0,\ j\geq1.
1089: \ea
1090: \]
1091: $I_{ij}$ is the time it takes for particle $P_j$ to jump again after its jump to site $i-j$.
1092: %Equivalently, it is the time between the departures of customers $i-1$ and $i$ from server $j$.
1093: $J_{ij}$ is the time it takes for hole $H_i$ to jump again after its jump to site $i-j+1$.
1094: %Equivalently, it is the time between the departures of customer $i$ from servers $j-1$ and $j$.
1095: Applying the last passage rules \eqref{eq:grec} shows
1096: \be
1097: \begin{split}
1098: I_{ij}&=G_{ij}-G_{\{i-1\}j}\\
1099: &=(G_{\{i-1\}j}\lor G_{i\{j-1\}})+\om_{ij}-G_{\{i-1\}\{j-1\}}-(G_{\{i-1\}j}-G_{\{i-1\}\{j-1\}})\label{eq:iij}\\
1100: %&=([G_{\{i-1\}j}-G_{\{i-1\}\{j-1\}}]\lor[G_{i\{j-1\}}-G_{\{i-1\}\{j-1\}}])+\om_{ij}-J_{\{i-1\}j}\\
1101: &=(J_{\{i-1\}j}\lor I_{i\{j-1\}})+\om_{ij}-J_{\{i-1\}j}\\
1102: &=(I_{i\{j-1\}}-J_{\{i-1\}j})^++\om_{ij}.
1103: \end{split}
1104: \ee
1105: Similarly,
1106: \be
1107: J_{ij}=(J_{\{i-1\}j}-I_{i\{j-1\}})^++\om_{ij}.\label{eq:jij}
1108: \ee
1109: For later use, we define
1110: \be
1111: X_{\{i-1\}\{j-1\}}=I_{i\{j-1\}}\land J_{\{i-1\}j}.\label{eq:xdef}
1112: \ee
1113: \begin{lm}\label{lm:markov}
1114: Fix $i,\,j\geq1$. If $I_{i\{j-1\}}$ and $J_{\{i-1\}j}$ are independent
1115: exponentials with respective parameters $1-\vr$ and $\vr$, then $I_{ij}$, $J_{ij}$, and $X_{\{i-1\}\{j-1\}}$ are jointly independent exponentials with respective parameters $1-\vr$, $\vr$, and $1$.
1116: \end{lm}
1117: \begin{proof}
1118: As the variables $I_{i\{j-1\}}$, $J_{\{i-1\}j}$ and $\om_{ij}$ are independent, we use \eqref{eq:iij}, \eqref{eq:jij} and \eqref{eq:xdef} to write the joint moment generating function as
1119: \begin{multline*}
1120: M_{I_{ij},\,J_{ij},\,X_{\{i-1\}\{j-1\}}}(s,\,t,\,u):=\Ev\e{sI_{ij}+tJ_{ij}+uX_{\{i-1\}\{j-1\}}}\\
1121: =\Ev\e{s(I_{i\{j-1\}}-J_{\{i-1\}j})^++t(J_{\{i-1\}j}-I_{i\{j-1\}})^++u(I_{i\{j-1\}}\land J_{\{i-1\}j})}\cdot\Ev\e{(s+t)\om_{ij}}
1122: \end{multline*}
1123: where it is defined. Then, with the assumption of the lemma
1124: and the definition of $\om_{ij}$, elementary calculations show
1125: \[
1126: M_{I_{ij},\,J_{ij},\,X_{\{i-1\}\{j-1\}}}(s,\,t,\,u)=\frac{\vr\cdot(1-\vr)}{(1-\vr-s)\cdot(\vr-t)\cdot(1-u)}.
1127: \]
1128: \end{proof}
1129: 
1130: 
1131: 
1132: 
1133: 
1134: 
1135: Let $\Sigma$ be the set of doubly-infinite down-right paths
1136: in the first quadrant of the $(i,j)$-coordinate system.
1137:  In terms of the sequence of points visited
1138:  a path $\si\in\Sigma$ is given by
1139: \[
1140: \si=\{\dots\to
1141:  (p_{-1}, \,q_{-1})\to (p_0,\,q_0)\to(p_1,\,q_1)\to\dots
1142: \to(p_l,\,q_l)\to\dots\}
1143: \]
1144: with all $p_l,q_l\geq 0$ and
1145:  steps
1146: \[
1147: (p_{l+1},\,q_{l+1})-(p_l,\,q_l)= %%%(1,0)\ \text{ or }\ (0,-1)
1148: \begin{cases} (1,0) &\text{(direction $\to$ in Figure \ref{fig:iniri}), or}\\
1149: (0,-1) &\text{(direction $\downarrow$  in Figure \ref{fig:iniri}).}
1150: \end{cases}
1151: \]
1152: The interior
1153: of the set enclosed by $\si$ is defined by
1154: \[
1155: \mathcal B(\si)=\{(i,\,j)\,:\,0\leq i<p_l,\ 0\leq j<q_l\text{ for some }(p_l,\,q_l)\in\si\}.
1156: \]
1157: The last-passage time increments along $\si$ are the variables
1158: \[
1159:  Z_l(\si)=G_{p_{l+1}q_{l+1}}-G_{p_lq_l}=
1160:  \left\{\ba
1161: &I_{p_{l+1}q_{l+1}},&&\text{if }(p_{l+1},\,q_{l+1})-(p_l,\,q_l)=(1,0),\\
1162: &J_{p_lq_l},&&\text{if }(p_{l+1},\,q_{l+1})-(p_l,\,q_l)=(0,-1),
1163: \ea\right.
1164: \]
1165: for $l\in\Zb$.
1166: We admit the possibility that  $\si$ is the union of the $i$- and
1167: $j$-coordinate
1168: axes, in which case $\mathcal B(\si)$ is empty.
1169: \begin{lm}\label{lm:NE}
1170: For any $\si\in\Sigma$,  the random variables
1171: \be
1172: \bigl\{\{X_{ij}\,:\,(i,\,j)\in\mathcal B(\si)\},\  \{Z_l(\si): l\in\Zb\}
1173: \bigr\}\label{eq:mind}
1174: \ee
1175: are mutually independent, $I$'s with $\Expd(1-\vr)$, $J$'s with
1176:  $\Expd(\vr)$, and $X$'s with $\Expd(1)$ distribution.
1177: \end{lm}
1178: \begin{proof}  We first consider the countable set of
1179: paths that join the $j$-axis to the $i$-axis, in other words
1180: those for which there exist
1181:  finite $n_0<n_1$ such that
1182: $p_{n}=0$ for $n\leq n_0$ and  $q_{n}=0$ for $n\geq n_1$.
1183: For these paths we argue by induction on $\mathcal B(\si)$.
1184: When $\mathcal B(\si)$ is the empty set,
1185:  the statement reduces to the independence of $\om$-values on the $i$- and $j$-axes
1186: which is part of the set-up.
1187: 
1188: Now given an arbitrary $\si\in\Sigma$ that connects the $j$- and the $i$-axes, consider
1189: a {\sl growth corner} $(i,j)$ for $\mathcal B(\si)$,
1190: by which we mean that for some index $l\in\Zb$,
1191: \[ (p_{l-1}, q_{l-1}), (p_{l}, q_{l}), (p_{l+1}, q_{l+1})
1192: \;=\;
1193: (i,j+1), (i,j), (i+1,j). \]
1194: A new valid $\wt\si\in\Sigma$ can be produced by
1195: replacing the above points with
1196: \[ (\wt p_{l-1}, \wt q_{l-1}), (\wt p_{l}, \wt q_{l}),
1197: (\wt p_{l+1}, \wt q_{l+1})
1198: \;=\;
1199: (i,j+1), (i+1,j+1), (i+1,j) \]
1200: and now $\mathcal B(\wt\si)=\mathcal B(\si)\cup\{(i,j)\}$.
1201: 
1202: The change inflicted on the set of random variables
1203: \eqref{eq:mind} is that
1204: \be
1205: \{I_{\{i+1\}j},\ J_{i\{j+1\}}\}\label{eq:rem}
1206: \ee
1207: has been replaced by
1208: \be
1209: \{I_{\{i+1\}\{j+1\}},\ J_{\{i+1\}\{j+1\}},\ X_{ij}\}.\label{eq:add}
1210: \ee
1211: By \eqref{eq:iij}--\eqref{eq:jij}
1212: variables  \eqref{eq:add} are determined by \eqref{eq:rem} and
1213: $\om_{\{i+1\}\{j+1\}}$.  If we assume
1214: inductively that  $\si$ satisfies the conclusion we seek, then
1215: so does $\wt\si$ by
1216:  Lemma \ref{lm:markov} and because in the situation under
1217: consideration  $\om_{\{i+1\}\{j+1\}}$ is
1218: independent of the variables in \eqref{eq:mind}.
1219: 
1220:  For an arbitrary $\sigma$ the statement follows
1221: because the independence of the random variables in \eqref{eq:mind}
1222: follows from independence of finite subcollections.
1223: Consider any square $R=\{0\leq i,j\leq M\}$  large enough so that
1224: the corner $(M,M)$ lies outside $\si\cup\mathcal B(\si)$.
1225: Then the $X$- and $Z(\sigma)$-variables associated to $\si$
1226: that lie in $R$ are a subset of the variables of a certain
1227: path $\wt\si$ that goes through the points $(0,M)$ and $(M,0)$.
1228: Thus the variables
1229: in \eqref{eq:mind} that lie inside an arbitrarily large square
1230: are independent.
1231: \end{proof}
1232: 
1233: 
1234: %In particular, this lemma states that two straight last-passage paths either leaving a common point or arriving to a common point meet jointly independent $G$-increments. Compare this result to the first paragraph of the proof of Theorem 2.1 in
1235: % Cator and Groeneboom \cite{cuberoot}.
1236: 
1237: By applying Lemma \ref{lm:NE}
1238: to a path that contains the horizontal line  $q_l\equiv j$ we get
1239: a version of Burke's theorem: particle
1240: $P_j$ obeys a Poisson process after time
1241: $G_{0j}$ when it  ``enters the last-passage picture.''
1242: The vertical line  $p_l\equiv i$
1243: gives the corresponding  statement for hole $H_i$.
1244: 
1245: Example 2.10.2 of Walrand \cite{walrand} gives an
1246: intuitive understanding of this result. Our initial state
1247: corresponds to the situation when particle $P_0$
1248: and hole $H_0$ have just exchanged places
1249:  in an equilibrium system of queues. $H_0$ is
1250: therefore a customer who has just moved from queue 0 to queue 1.
1251: By that Example, this customer sees an equilibrium system of
1252: queues every time he jumps. Similarly, any new customer arriving
1253: to the queue of particle $P_1$ sees an equilibrium queue system in
1254: front, so Burke's theorem extends to the region between $P_0$ and $H_0$.
1255: 
1256: Up-right turns do not have independence:
1257: variables $I_{ij}$ and $J_{i\{j+1\}}$, or $J_{ij}$ and $I_{\{i+1\}j}$
1258: are not independent.
1259: %However, ``light turns'', i.e.\ paths consisting of steps $\to$ and $\searrow$ (this second kind going against last-passage allowable steps) will meet independent $G$-increments.
1260: 
1261: The same inductive argument with a growing cluster $\mathcal B(\si)$
1262:  proves a result that corresponds to a coupling of
1263:  two exclusion systems $\eta$ and $\wt\eta$ where the
1264: latter has a higher density of particles.
1265: However, the lemma is a purely deterministic
1266: statement.
1267: 
1268: \begin{lm}\label{lm:basic1}
1269: Consider two assignments of values $\{\om_{ij}\}$ and $\{\wt\om_{ij}\}$
1270: that satisfy $\om_{00}=\wt\om_{00}=0$,
1271:  $\om_{0j}\geq\wt\om_{0j}$, $\om_{i0}\leq\wt\om_{i0}$,
1272: and $\om_{ij}=\wt\om_{ij}$  for all $i,j\geq 1$.     Then all
1273: increments satisfy  $I_{ij}\leq\wt I_{ij}$ and  $J_{ij}\geq\wt J_{ij}$.
1274: \end{lm}
1275: \begin{proof}  One  proves by induction that the statement holds for
1276: all increments between points in $\si\cup\mathcal B(\si)$ for
1277: those paths $\si\in\Sigma$ for which $\mathcal B(\si)$ is finite.
1278: If $\mathcal B(\si)$ is empty the statement is the assumption
1279: made on the $\om$- and $\wt\om$-values on the $i$- and $j$-axes. The induction
1280: step that adds a growth corner to $\mathcal B(\si)$
1281:  follows from equations \eqref{eq:iij} and \eqref{eq:jij}.
1282: \end{proof}
1283: 
1284: \subsection{The reversed process.}
1285: 
1286: Fix $m>0$ and $n>0$, and define
1287: \[
1288: H_{ij}=G_{mn}-G_{ij}
1289: \]
1290: for $0\leq i\leq m$, $0\leq j\leq n$. This is the time needed to ``free'' the
1291: point $(i,\,j)$ in the reversed process, started from the moment when $(m,\,n)$
1292: becomes occupied.  For $0\leq i<m$ and $0\leq j<n$,
1293: \[
1294: \ba
1295: H_{ij}&%=G_{mn}-G_{ij}\\
1296: %&=G_{mn}-(G_{\{i+1\}j}\land G_{i\{j+1\}})+(G_{\{i+1\}j}\land G_{i\{j+1\}})-G_{ij}\\
1297: =-((G_{\{i+1\}j}-G_{mn})\land(G_{i\{j+1\}}-G_{mn}))\\
1298: &\qquad+((G_{\{i+1\}j}-G_{ij})\land(G_{i\{j+1\}}-G_{ij}))\\
1299: %&=((G_{mn}-G_{\{i+1\}j})\lor(G_{mn}-G_{i\{j+1\}}))\\
1300: %&\qquad+((G_{\{i+1\}j}-G_{ij})\land(G_{i\{j+1\}}-G_{ij}))\\
1301: %&=H_{\{i+1\}j}\lor H_{i\{j+1\}}+I_{\{i+1\}j}\land J_{i\{j+1\}}\\
1302: &=H_{\{i+1\}j}\lor H_{i\{j+1\}}+X_{ij}
1303: \ea
1304: \]
1305: with definition \eqref{eq:xdef} of the $X$-variables. Taking this and Lemma \ref{lm:NE} into account, we see that the $H$-process is a
1306: copy of the original $G$-process, but with reversed coordinate directions.
1307: Precisely speaking,  define $\om^*_{00}=0$,
1308: and then for $0<i\leq m$, $0<j\leq n$: $\om^*_{i0}=I_{\{m-i+1\}n}$, $\om^*_{0j}=J_{m\{n-j+1\}}$,
1309: and $\om^*_{ij}=X_{\{m-i\}\{n-j\}}$.  Then  $\{\om^*_{ij}: 0\leq i\leq m, 0\leq j\leq n\}$
1310: is distributed like $\{\om_{ij}: 0\leq i\leq m, 0\leq j\leq n\}$ in \eqref{eq:bondis}, and
1311:  the process
1312: \be
1313: G^*_{ij}=H_{\{m-i\}\{n-j\}}\label{eq:def-K}
1314: \ee
1315: for $0\leq i\leq m$, $0\leq j\leq n$ satisfies
1316: \[
1317: G^*_{ij}=(G^*_{\{i-1\}j}\lor G^*_{i\{j-1\}})+\om^*_{ij}, \qquad 0\leq i\leq m\,,\, 0\leq j\leq n
1318: \]
1319: (with the formal assumption $G^*_{\{-1\}j}=G^*_{i\{-1\}}=0$), see \eqref{eq:grec}.
1320: Thus the pair  $(G^*,\om^*)$ has the same distribution as $(G,\om)$ in a fixed
1321: rectangle $\{0\leq i\leq m\}\times \{ 0\leq j\leq n\}$. Throughout the paper quantities defined in the reversed process will be denoted by a superscript $^*$, and they will always be equal in distribution to their original forward versions.
1322: 
1323: \subsection{Exit point and competition interface.}
1324: 
1325: For integers $x$ define
1326: \be
1327: U^\vr_x=G_{\{x^+\}\{x^-\}}=\begin{cases}
1328: \sum_{i=0}^x \om_{i0}, &x\geq 0\\
1329: \sum_{j=0}^{-x} \om_{0j}, &x\leq 0.
1330: \end{cases} \label{eq:udef}
1331: \ee
1332:  Referring to the two
1333: coordinate systems in Figure \ref{fig:ini},
1334: this is  the last-passage time of
1335:  the point on the $(i,j)$-axes  above point  $x$ on the $x$-axis.
1336: This point is on the $i$-axis
1337: if $x\geq0$ and  on the $j$-axis if $x\leq0$.
1338: 
1339: Fix  integers $m\geq x^+\lor 1$, $n\geq x^-\lor 1$, and define $\Pi_x(m,n)$
1340: as the set of directed paths $\pi$ connecting
1341: $(x^+\lor 1,\,x^-\lor 1)$ and $(m,\,n)$
1342: using allowable steps \eqref{eq:allow}. Then let
1343: \be
1344: A_x=A_x(m,n)=\max_{\pi\in\Pi_x(m,n)}\sum_{(p,q)\in\pi}\om_{pq}\label{eq:adef}
1345: \ee
1346: be the maximal weight collected by a path from
1347: $(x^+,\,x^-)$ to $(m,\,n)$ that immediately exits the axes,
1348: and does not  count $\om_{x^+x^-}$.
1349: Notice that $A_{-1}=A_0=A_1$ and this value is the last-passage
1350: time from $(1,1)$ to $(m,n)$ that completely ignores the
1351: boundaries, or in other words, sets the boundary values
1352: $\om_{i0}$ and $\om_{0j}$  equal
1353: to zero.
1354: 
1355: By the continuity of the exponential distribution
1356:  there is an a.s.\ unique path $\wh\pi$
1357:  from $(0,\,0)$ to $(m,\,n)$ which
1358: collects the maximal weight $G^\vr_{mn}$.   Earlier we defined
1359: the {\sl exit point}  $Z^\vr\in\Zb$ to represent  the last
1360: point of this path on either the $i$-axis or the $j$-axis.
1361: Equivalently we can now state that
1362:  $Z^\vr$ is the a.s.\ unique integer for which
1363: \[
1364: G_{mn}^\vr=U^\vr_{Z^\vr}+A_{Z^\vr}.
1365: \]
1366: Simply because the maximal path $\wh\pi$ necessarily goes through either
1367: $(0,1)$ or $(1,0)$, $Z^\vr$ is always nonzero.
1368: 
1369: %%\hop
1370: 
1371: Recall the definition of the competition interface in \eqref{eq:def-vphi}.
1372: Now we can observe
1373: that the competition interface is the time reversal
1374: of the maximal path $\wh\pi$. Namely, the competition interface
1375: of the reversed process follows the maximal path $\wh\pi$ backwards
1376: from the corner  $(m,n)$, until it hits either the $i$- or the
1377: $j$-axis.  To make a precise statement, let us represent the
1378: a.s.\ unique maximal last-passage
1379: path, with exit point $Z^\vr$ as defined above, as
1380: \[
1381: \wh\pi=\{ (0,0)=\wh\pi_0 \to \wh\pi_1 \to \dotsm \to \wh\pi_{\lvert Z^\vr\rvert}
1382: \to\dots \to \wh\pi_{m+n}=(m,n)\},
1383: \]
1384: where  $\{\wh\pi_{\lvert Z^\vr\rvert+1}
1385: \to\dots \to \wh\pi_{m+n}\}$ is the portion of the path
1386: that  resides in the interior
1387: $\{1,\dotsc,m\}\times\{1,\dotsc,n\}$.
1388: 
1389: \begin{lm}
1390:   Let $\vp^*$
1391: be the competition interface constructed for the process
1392: $G^*$ defined by \rm{\eqref{eq:def-K}}. Then
1393: $\vp_k^*=(m,n)-\wh\pi_{m+n-k}$ for $0\leq k\leq m+n-\lvert Z^\vr\rvert$.
1394: \end{lm}
1395: \begin{proof}  Starting from $\wh\pi_{m+n}=(m,n)$,
1396: the maximal path $\wh\pi$ can be constructed backwards
1397: step by step  by always moving to the maximizing point
1398: of the right-hand side of \eqref{eq:grec}.
1399: % the equation
1400: %\[
1401: %G_{ij}= G_{\{i-1\}j}\lor G_{i\{j-1\}}+\om_{ij}.
1402: %\]
1403: This is the same as constructing the competition interface
1404: for the reversed process $G^*$ by \eqref{eq:def-vphi}. Since $G^*$ is not constructed
1405: outside the rectangle $\{0,\dotsc,m\}\times\{0,\dotsc,n\}$, we
1406: cannot assert what the competition interface does after the
1407: point
1408: \[\vp^*_{m+n-\lvert Z^\vr\rvert}=
1409: \wh\pi_{\lvert Z^\vr\rvert}=\begin{cases} (Z^\vr, 0) &\text{if $Z^\vr>0$}\\
1410: (0,-Z^\vr) &\text{if $Z^\vr<0$.}
1411: \end{cases}
1412: \qedhere
1413: \]
1414: \end{proof}
1415: Notice that, due to this lemma, $Z^{*\vr}$ defined in \eqref{eq:zstar} is indeed $Z^\vr$ defined in the reversed process, which justifies the argument following \eqref{eq:zstar}.
1416: %Recall \eqref{eq:def-vn} and \eqref{eq:def-wm}, and notice also that in the situation of the lemma,
1417: %\be
1418: %\begin{split}
1419: %&\text{$Z^\vr>0 \;\Longleftrightarrow\; v(n)<m$, and then
1420: % $v(n)=m-Z^\vr$, while}\\
1421: %&\text{$Z^\vr<0 \;\Longleftrightarrow\; w(m)<n$, and then  $w(m)=n+Z^\vr$. }
1422: %\end{split}
1423: %\label{eq:Z-vphi}
1424: %\ee
1425: 
1426: The competition interface
1427: bounds  the regions  where the boundary conditions on the  axes are
1428: felt.  From this we can get useful
1429:  bounds between last-passage times under different
1430: boundary conditions.  This is  the last-passage model equivalent
1431: of the common use of second-class particles to control discrepancies
1432: between coupled  interacting particle systems. In the next lemma,
1433: the superscript $\Wc$ represents the west boundary ($j$-axis) of the
1434: $(i,j)$-plane. Remember that $(v(n),n)$ is the left-most point of the competition interface on the horizontal line $j=n$
1435: computed in terms of the $G$-process (see \eqref{eq:def-vn}).
1436: 
1437: \begin{lm} Let $G^{\Wc=0}$ be the last-passage times of a system where
1438: we set $\om_{0j}=0$ for all $j\geq 1$. Then for $v(n)<m_1<m_2$,
1439: \begin{align*}
1440: A_0(m_2,n)-A_0(m_1,n)&\leq G^{\Wc=0}(m_2,n)-G^{\Wc=0}(m_1,n)\\
1441: &= G(m_2,n)-G(m_1,n).
1442: \end{align*}
1443: \label{lm:coup-3}
1444: \end{lm}
1445: \begin{proof} The first inequality is a consequence of
1446: Lemma \ref{lm:basic1}, because computing $A_0$ is the same
1447: as computing $G$ with all boundary values $\om_{i0}=\om_{0j}=0$ (and in fact this inequality is valid for all $m_1<m_2$.)
1448: The equality $G(m,n)=G^{\Wc=0}(m,n)$ for $m>v(n)$ follows because
1449: the maximal path $\wh\pi$ for $G(m,n)$ goes through $(1,0)$ and
1450: hence does not see the boundary values $\om_{0j}$. Thus this
1451: same path $\wh\pi$ is maximal for $G^{\Wc=0}(m,n)$ too.
1452: \end{proof}
1453: 
1454: If we set $\om_{i0}=0$ (the south boundary denoted by $\Sc$)
1455: instead, we get this
1456: statement: for $0\leq m_1<m_2\leq v(n)$,
1457: \be
1458: \begin{split}
1459: A_0(m_2,n)-A_0(m_1,n)&\geq G^{\Sc=0}(m_2,n)-G^{\Sc=0}(m_1,n)\\
1460: &= G(m_2,n)-G(m_1,n).
1461: \label{eq:coup-4}
1462: \end{split}
1463: \ee
1464: 
1465: \subsection{A coupling on the $i$-axis}
1466: 
1467: Let $1>\la>\vr>0$. As a common realization of the exponential weights $\om_{i0}^\la$ of $\Expd(1-\la)$ and $\om_{i0}^\vr$ of $\Expd(1-\vr)$ distribution, we write
1468: \be
1469: \om_{i0}^\la=\frac{1-\vr}{1-\la}\cdot\om_{i0}^\vr.\label{eq:cou}
1470: \ee
1471: We will use this coupling later for different purposes. We will also need
1472: \[
1473: \Vv(\om_{i0}^\la-\om_{i0}^\vr)=\left(\frac{1-\vr}{1-\la}-1\right)^2\cdot\frac{1}{(1-\vr)^2}=\left(\frac{1}{1-\la}-\frac{1}{1-\vr}\right)^2.
1474: \]
1475: 
1476: \subsection{Exit point and the variance of the last-passage time}
1477: 
1478: With these preliminaries
1479: we can prove  the key lemma that links the variance of the last-passage
1480: time to the weight collected along the axes.
1481: 
1482: 
1483: \begin{lm}\label{lm:ns}
1484: Fix $m,\,n$ positive integers. Then
1485: \be
1486: \ba
1487: \Vv(G^\vr_{mn})&=\frac{n}{\vr^2}-\frac{m}{(1-\vr)^2}+\frac{2}{1-\vr}\cdot\Ev(U_{Z^{\vr+}}^\vr)\\
1488: &=\frac{m}{(1-\vr)^2}-\frac{n}{\vr^2}+\frac{2}{\vr}\cdot\Ev(U_{-Z^{\vr-}}^\vr),
1489: \ea\label{eq:ns}
1490: \ee
1491: where $Z^\vr$ is the a.s.\ unique exit point of the maximal path from $(0,\,0)$ to $(m,\,n)$.
1492: \end{lm}
1493: \begin{proof}
1494: We label the total increments along the sides of the rectangle by compass
1495: directions:
1496: \[
1497: \Wc=G_{0n}^\vr-G_{00}^\vr,\quad\Nc=G_{mn}^\vr-G_{0n}^\vr,\quad\Ec=G_{mn}^\vr-G_{m0}^\vr,\quad\Sc=G_{m0}^\vr-G_{00}^\vr.
1498: \]
1499: As $\Nc$ and $\Ec$ are independent by Lemma \ref{lm:NE}, we have
1500: \be
1501: \ba
1502: \Vv(G^\vr_{mn})&=\Vv(\Wc+\Nc)\\
1503: &=\Vv(\Wc)+\Vv(\Nc)+2\Cov(\Sc+\Ec-\Nc,\,\Nc)\\
1504: &=\Vv(\Wc)-\Vv(\Nc)+2\Cov(\Sc,\,\Nc).
1505: \ea\label{eq:varg}
1506: \ee
1507: We now modify the $\om$-values in $\Sc$. Let $\la=\vr+\ve$ and
1508: apply \eqref{eq:cou}, without changing the other values
1509:  $\{\om_{ij}: i\geq 0,j\geq1\}$.  Quantities of the altered last-passage
1510:  model will be marked with a superscript $\ve$. In
1511: this new process, $\Sc^\ve$ has a Gamma$(m,\,1-\vr-\ve)$ distribution with density
1512: \[
1513: f_\ve(s)=\frac{(1-\vr-\ve)^m\cdot\e{-(1-\vr-\ve)s}\cdot s^{m-1}}{(m-1)!}
1514: \]
1515: for $s>0$, whose $\ve$-derivative is
1516: \be
1517: \pt_\ve f_\ve(s)=sf_\ve(s)-\frac{m}{1-\vr-\ve}\cdot f_\ve(s).\label{eq:fe}
1518: \ee
1519: Given the sum $\Sc^\ve$, the joint distribution of $\{\om_{i0}\}_{1\leq i\leq m}$ is independent of the parameter $\ve$, hence the quantity $\Ev(\Nc^\ve\,|\,\Sc^\ve=s)=\Ev(\Nc\,|\,\Sc=s)$ does not depend on $\ve$. Therefore, using \eqref{eq:fe} we have
1520: \be
1521: \begin{split}
1522: \pt_\ve\Ev(\Nc^\ve)\Bigr|_{\ve=0}&=\pt_\ve\int_0^\infty\Ev(\Nc\,|\,\Sc=s)f_\ve(s)\,\di s\Bigr|_{\ve=0}\\
1523: &=\int_0^\infty\Ev(\Nc\,|\,\Sc=s)\cdot s\cdot f_0(s)\,\di s-\frac{m}{1-\vr}\int_0^\infty\Ev(\Nc\,|\,\Sc=s)f_0(s)\,\di s\\
1524: &=\Ev(\Nc\Sc)-\frac{m}{1-\vr}\cdot\Ev(\Nc)=\Cov(\Nc,\,\Sc).\label{eq:dcov}
1525: \end{split}
1526: \ee
1527: Next we compute the same quantity by a different approach. Let $Z$ and $Z^\ve$ be the exit points of the maximal paths to $(m,\,n)$ in the original and the modified processes, respectively. Similarly, $U_x$ and $U^\ve_x$ are the weights as defined by \eqref{eq:udef} for the two processes. Hence $U_Z$ is the weight collected on the $i$ or $j$ axis by the maximal path of the original process. Then
1528: \[
1529: \ba
1530: \Nc^\ve-\Nc&=(\Nc^\ve-\Nc)\cdot{\bf1}\{Z^\ve=Z\}+(\Nc^\ve-\Nc)\cdot{\bf1}\{Z^\ve\neq Z\}\\
1531: &=(U_Z^\ve-U_Z)\cdot{\bf1}\{Z^\ve=Z\}+(\Nc^\ve-\Nc)\cdot{\bf1}\{Z^\ve\neq Z\}\\
1532: &=(U_Z^\ve-U_Z)+(\Nc^\ve-\Nc-U_Z^\ve+U_Z)\cdot{\bf1}\{Z^\ve\neq Z\}.
1533: \ea
1534: \]
1535: As $\om$ values are only changed on the $i$-axis, the first term is rewritten as
1536: \[
1537: U_Z^\ve-U_Z=U_{Z^+}^\ve-U_{Z^+}=\left(\frac{1-\vr}{1-\vr-\ve}-1\right)U_{Z^+}=\frac{\ve}{1-\vr-\ve}\cdot U_{Z^+}
1538: \]
1539: by \eqref{eq:cou}. We show that the expectation
1540: of the second term is $\mathfrak o(\ve)$.   Note that the
1541: increase $\Nc^\ve-\Nc$ is bounded by $\Sc^\ve-\Sc$. Hence
1542: \be
1543: \begin{split}
1544: &\Ev[(\Nc^\ve-\Nc-U_Z^\ve+U_Z)\cdot{\bf1}\{Z^\ve\neq Z\}]\\
1545: &\qquad \leq\Ev[(\Nc^\ve-\Nc)\cdot{\bf1}\{Z^\ve\neq Z\}]
1546: \;\leq\;\Ev[(\Sc^\ve-\Sc)\cdot{\bf1}\{Z^\ve\neq Z\}]\\
1547: &\qquad \leq\bigl(\Ev[(\Sc^\ve-\Sc)^2]\bigr)^\frac12\cdot
1548: \bigl(\Pv\{Z^\ve\neq Z\}\bigr)^\frac12.
1549: \end{split}\label{eq:err}
1550: \ee
1551: To show that the probability is of the order of $\ve$, notice that the exit point of the maximal path can only differ in the modified process from the one of the original process, if for some $Z<k\leq m$, $U^\ve_k+A_k>U^\ve_Z+A_Z$ with $Z$ of the original process (see \eqref{eq:adef} for the definition of $A_i$). Therefore,
1552: \[
1553: \ba
1554: \Pv\{Z^\ve\neq Z\}&=\Pv\{U^\ve_k+A_k>U^\ve_Z+A_Z\text{ for some }Z<k\leq m\}\\
1555: &=\Pv\{U^\ve_k-U^\ve_Z>A_Z-A_k\text{ for some }Z<k\leq m\}\\
1556: &=\Pv\{U^\ve_k-U^\ve_Z>A_Z-A_k\geq U_k-U_Z\text{ for some }Z<k\leq m\}\\
1557: &\leq\Pv\{U^\ve_k-U^\ve_i>A_i-A_k\geq U_k-U_i\text{ for some }0\leq i<k\leq m\}\\
1558: &\leq\sum_{0\leq i<k\leq m}\Pv\{U^\ve_k-U^\ve_i>A_i-A_k\geq U_k-U_i\}.
1559: \ea
1560: \]
1561: We also used the definition of $Z$ in the third equality, via $A_Z+U_Z\geq A_k+U_k$. Notice that $A$'s and $U$'s are independent for fixed indices. Hence with $\mu$ denoting the distribution of $A_i-A_k$, we write
1562: \begin{multline*}
1563: \Pv\{U^\ve_k-U^\ve_i>A_i-A_k\geq U_k-U_i\}\\
1564: \ba
1565: &=\int\Pv\{U^\ve_k-U^\ve_i>x\geq U_k-U_i\}\di\mu(x)\\
1566: &\leq\sup_x\Pv\{U^\ve_k-U^\ve_i>x\geq U_k-U_i\}\\
1567: &=\sup_x\Pv\Bigl\{\frac{1-\vr}{1-\vr-\ve}\cdot(U_k-U_i)>x\geq U_k-U_i\Bigr\}\\
1568: &=\sup_x\Pv\Bigl\{x\geq U_k-U_i>x\Bigl(1-\frac{\ve}{1-\vr}\Bigr)\Bigr\}.
1569: \ea
1570: \end{multline*}
1571: Since $U_k-U_i$ has a Gamma distribution, the supremum above is $\mathcal O(\ve)$, which shows the bound on $\Pv\{Z^\ve\neq Z\}$. The first factor on the right-hand side of \eqref{eq:err},
1572: \[
1573: \bigl(\Ev[(\Sc^\ve-\Sc)^2]\bigr)^{1/2}
1574: =\frac{\ve}{1-\vr-\ve}\cdot\bigl(\Ev[\Sc^2]\bigr)^{1/2},
1575: \]
1576: is of order $\ve$.
1577: Hence the error term \eqref{eq:err} is $\mathfrak o(\ve)$, and we conclude
1578: \[
1579: \pt_\ve\Ev(\Nc^\ve)\Bigr|_{\ve=0}=\frac{1}{1-\vr}\cdot\Ev(U_{Z^+}).
1580: \]
1581: The proof of the first statement is then completed by this display, \eqref{eq:varg} and \eqref{eq:dcov}, as $\Wc$ and $\Nc$ are Gamma-distributed by Lemma \ref{lm:NE}. The second statement follows in a similar way, using $\Cov(\Wc,\,\Ec)$.
1582: \end{proof}
1583: 
1584: \begin{lm}\label{lm:vvlr}
1585: Let $0<\vr\leq\la<1$. Then
1586: \[
1587: \Vv(G_{mn}^\la)\leq\frac{\vr^2}{\la^2}\cdot\Vv(G^\vr_{mn})+m\cdot\left(\frac{1}{(1-\la)^2}-\frac{\vr^2}{\la^2(1-\vr)^2}\right).
1588: \]
1589: \end{lm}
1590: \begin{proof}
1591: The proof is based on the coupling described by \eqref{eq:cou}, and a similar one $\om_{0j}^\la=\frac\vr\la\cdot\om_{0j}^\vr$ on the $j$ axis. Note that in this coupling, when changing from $\vr$ to $\la$, we are increasing the weights on the $i$-axis and decreasing the weights on the $j$-axis, which clearly implies $Z^\vr\leq Z^\la$. Also, we remain in the stationary situation, so \eqref{eq:ns} remains valid for $\la$. As $U_{-x^-}^\vr$ is non-increasing in $x$, this implies
1592: \[
1593: U^\la_{-Z^{\la-}}=\frac\vr\la\cdot U^\vr_{-Z^{\la-}}\leq\frac\vr\la\cdot U^\vr_{-Z^{\vr-}}.
1594: \]
1595: We substitute this into the second line of \eqref{eq:ns} to get
1596: \[
1597: \ba
1598: \Vv(G^\la_{mn})&=\frac{m}{(1-\la)^2}-\frac{n}{\la^2}+\frac{2}{\la}\cdot\Ev(U_{-Z^{\la-}}^\la)\\
1599: &\leq\frac{m}{(1-\la)^2}-\frac{n}{\la^2}+\frac{2\vr}{\la^2}\cdot\Ev(U_{-Z^{\vr-}}^\vr)\\
1600: &=\frac{\vr^2}{\la^2}\cdot\left(\frac{m}{(1-\vr)^2}-\frac{n}{\vr^2}+\frac{2}{\vr}\cdot\Ev(U_{-Z^{\vr-}}^\vr)\right)\\
1601: &\qquad+m\cdot\left(\frac{1}{(1-\la)^2}-\frac{\vr^2}{\la^2(1-\vr)^2}\right)\\
1602: &=\frac{\vr^2}{\la^2}\cdot\Vv(G^\vr_{mn})+m\cdot\left(\frac{1}{(1-\la)^2}-\frac{\vr^2}{\la^2(1-\vr)^2}\right).
1603: \ea
1604: \]
1605: \end{proof}
1606: 
1607: %%\hop
1608: %%{\bf Behavior on the characteristics.}
1609: 
1610: \section{Upper bound}
1611: \label{sc:ub}
1612: 
1613: We turn  to proving the upper bounds in Theorems \ref{tm:G-var-1} and \ref{tm:Z-1}.
1614: We have a fixed  density
1615: $\vr\in(0,1)$, and to study the last-passage times $G^\vr$
1616:  along the  characteristic,
1617: we define the
1618: dimensions of the last-passage rectangle as
1619: \be
1620: m(t)=\fl{(1-\vr)^2t}\qquad\text{and}\qquad n(t)=\fl{\vr^2t}\label{eq:mndef}
1621: \ee
1622: with a parameter $t\to\infty$.
1623: The quantities $A_x$, $Z$ and $G_{mn}$ connected to these indices
1624: are denoted by $A_x(t)$, $Z(t)$, $G(t)$.
1625: In the proofs we  need to consider different  boundary conditions
1626:  \eqref{eq:bondis}  with $\vr$ replaced by $\la$.
1627: This will be indicated by a superscript.
1628:  However,  the
1629: superscript $\la$ only changes the boundary conditions
1630: and not the dimensions  $m(t)$ and $n(t)$, always defined by
1631: \eqref{eq:mndef} with a fixed $\vr$.
1632:  Moreover,
1633: we apply the coupling \eqref{eq:cou} on the $i$-axis and $\om_{0j}^\la=\frac\vr\la\cdot\om_{0j}^\vr$ on the $j$-axis.
1634: The weights $\{\om_{ij}\}_{i,\,j\geq1}$ in the interior
1635: will not be affected by changes in boundary conditions,
1636: so in particular  $A_x(t)$ will not either. Since
1637: $G^\la(t)$ chooses the  maximal path,
1638: \[
1639: U^\la_z+A_z(t)\leq G^\la(t)
1640: \]
1641: for all $1\leq z\leq m(t)$
1642: and all densities $0<\la<1$.
1643: Consequently, for integers $u\geq0$ and densities $\la\geq\vr$,
1644: \be
1645: \ba
1646: \Pv\{Z^\vr(t)>u\}&=\Pv\{\exists z>u\,:\,U^\vr_z+A_z(t)=G^\vr(t)\}\\
1647: &\leq\Pv\{\exists z>u\,:\,U^\vr_z-U^\la_z+G^\la(t)\geq G^\vr(t)\}\\
1648: &=\Pv\{\exists z>u\,:\,U^\la_z-U^\vr_z\leq G^\la(t)-G^\vr(t)\}\\
1649: &\leq\Pv\{U^\la_u-U^\vr_u\leq G^\la(t)-G^\vr(t)\}.
1650: \ea\label{eq:pzu}
1651: \ee
1652: The last step is justified by $\la\geq\vr$ and the coupling
1653: \eqref{eq:cou}.
1654: Set
1655: \be
1656: \la_u=\frac{\vr}{\sqrt{(1-\vr)^2-{u}/{t}}+\vr}.\label{eq:lu}
1657: \ee
1658: This density maximizes
1659: \[
1660: \Ev(U^\la_u)-\Ev(G^\la(t))=\frac{u}{1-\la}
1661: -\frac{\fl{(1-\vr)^2t}}{1-\la}-\frac{\fl{\vr^2t}}{\la}
1662: \]
1663: if the integer parts are dropped. The expectation
1664: $\Ev(G^\la_{mn})$ is
1665: computed as  $\Ev(G^\la_{0n})+\Ev(G^\la_{mn}-G^\la_{0n})$
1666: with the help of Lemma \ref{lm:NE}.
1667:  Some useful identities for future computations:
1668: \be
1669: \la_u\geq\vr,\quad\frac{1}{\la_u}=1+\frac{\sqrt{(1-\vr)^2-{u}/{t}}}{\vr},\quad\frac{1}{1-\la_u}=1+\frac{\vr}{\sqrt{(1-\vr)^2-{u}/{t}}}.\label{eq:usef}
1670: \ee
1671: 
1672: \begin{lm}
1673: With  $0\leq u\leq (1-\vr)^2t$ and  $\la_u$ of \eqref{eq:lu},
1674: \begin{multline*}
1675: \Ev(U^{\la_u}_u-U^\vr_u-G^{\la_u}(t)+G^\vr(t))\\
1676: \geq
1677: \frac{t\vr}{1-\vr}\left((1-\vr)-\sqrt{(1-\vr)^2-{u}/{t}}\,\right)^2
1678: -\frac{u/t}{\vr(1-\vr)}.
1679: \end{multline*}
1680: \end{lm}
1681: \begin{proof}
1682: By Lemma \ref{lm:NE} and \eqref{eq:mndef}
1683: \begin{align*}
1684: &\Ev(U^{\la_u}_u-U^\vr_u-G^{\la_u}(t)+G^\vr(t))\\
1685: &\quad =\frac{u}{1-\la_u}-\frac{u}{1-\vr}-\frac{\fl{(1-\vr)^2t}}{1-\la_u}+
1686: \frac{\fl{(1-\vr)^2t}}{1-\vr}-\frac{\fl{\vr^2t}}{\la_u}+
1687: \frac{\fl{\vr^2t}}{\vr}.
1688: \end{align*}
1689: First we remove the integer parts.
1690: Since $\la_u\geq\vr$,
1691: \[
1692: -\frac{\fl{(1-\vr)^2t}}{1-\la_u}+
1693: \frac{\fl{(1-\vr)^2t}}{1-\vr} \geq
1694: -\frac{(1-\vr)^2t}{1-\la_u}+
1695: \frac{(1-\vr)^2t}{1-\vr}.
1696: \]
1697: For the other integer parts
1698: \begin{align*}
1699: -\frac{\fl{\vr^2t}}{\la_u}+
1700: \frac{\fl{\vr^2t}}{\vr} &\geq  -\frac{\vr^2t}{\la_u}+
1701: \frac{\vr^2t}{\vr}  -\frac1\vr+\frac1\la_u\\
1702: &=-\frac{\vr^2t}{\la_u}+
1703: \frac{\vr^2t}{\vr}-\,\frac{1-\vr-\sqrt{(1-\vr)^2-u/t}}{\vr}\\
1704: &\geq -\frac{\vr^2t}{\la_u}+
1705: \frac{\vr^2t}{\vr} -\frac{u/t}{\vr(1-\vr)}
1706: \end{align*}
1707: The last term above
1708: is  the last term of the bound in the statement of the lemma.
1709: It remains to check that after  the integer parts have been
1710: removed from the mean, the remaining quantity  equals the main
1711:  term of the bound.
1712: \begin{multline*}
1713: \frac{u}{1-\la_u}-\frac{u}{1-\vr}-\frac{(1-\vr)^2t}{1-\la_u}+\frac{(1-\vr)^2t}{1-\vr}-\frac{\vr^2t}{\la_u}+\frac{\vr^2t}{\vr}\\
1714: \ba
1715: %&=[u-(1-\vr)^2t]\cdot\left[\frac{1}{1-\la_u}-\frac{1}{1-\vr}\right]-\vr^2t\cdot\left[\frac{1}{\la_u}-\frac{1}{\vr}\right]\\
1716: &=[u-(1-\vr)^2t]\cdot\left[1+\frac{\vr}{\sqrt{(1-\vr)^2-{u}/{t}}}-\frac{1}{1-\vr}\right]\\
1717: &\qquad-\vr^2t\cdot\left[1+\frac{\sqrt{(1-\vr)^2-{u}/{t}}}{\vr}-\frac{1}{\vr}\right]\\
1718: %&=t[(1-\vr)^2-{u}/{t}]\cdot\left[\frac{\vr}{1-\vr}-\frac{\vr}{\sqrt{(1-\vr)^2-{u}/{t}}}\right]\\
1719: %&\qquad+\vr^2t\cdot\left[\frac{1-\vr}{\vr}-\frac{\sqrt{(1-\vr)^2-{u}/{t}}}{\vr}\right]\\
1720: %&=t\cdot\frac{\vr}{1-\vr}\cdot\left[\left((1-\vr)^2-{u}/{t}\right)-2(1-\vr)\sqrt{(1-\vr)^2-{u}/{t}}+(1-\vr)^2\right]\\
1721: &=t\cdot\frac{\vr}{1-\vr}\left((1-\vr)-\sqrt{(1-\vr)^2-{u}/{t}}\right)^2.
1722: \ea
1723: \end{multline*}
1724: \end{proof}
1725: 
1726: \begin{lm}\label{lm:middle}
1727: For any $8\vr^{-2}(1-\vr)^2\leq u\leq(1-\vr)^2t$,
1728: \[
1729: \Ev(U^{\la_u}_u-U^\vr_u-G^{\la_u}(t)+G^\vr(t))\geq
1730: \frac{\vr}{8(1-\vr)^3}\cdot\frac{u^2}{t}.
1731: \]
1732: \end{lm}
1733: \begin{proof}
1734: Assumption $u\geq 8\vr^{-2}(1-\vr)^2$ implies that the last term
1735: of the bound from the previous lemma satisfies
1736: \[
1737: -\,\frac{u/t}{\vr(1-\vr)}\geq -\,\frac{\vr}{8(1-\vr)^3}\cdot\frac{u^2}{t}.
1738: \]
1739: Thus it remains to prove
1740: \[
1741: \left((1-\vr)-\sqrt{(1-\vr)^2-{u}/{t}}\,\right)^2\geq\frac{1}{4(1-\vr)^2}\cdot\frac{u^2}{t^2}.
1742: \]
1743: This is easy to check in the form
1744: \[
1745: %\ba
1746: \left(C-\sqrt{C^2-x}\right)^2%&
1747: \geq\frac{1}{4C^2}\cdot x^2%\\
1748: ,
1749: %C-\frac{1}{2C}\cdot x&\geq\sqrt{C^2-x}\\
1750: %C^2-x+\frac{1}{4C^2}\cdot x^2&\geq C^2-x\\
1751: %\frac{1}{4C^2}&\geq0,
1752: %\ea
1753: \]
1754: where  $x=u/t$, $C=1-\vr$ and then $x\leq C^2$.
1755: \end{proof}
1756: 
1757: \begin{lm}\label{lm:vg}
1758: For any $0\leq u\leq\frac34(1-\vr)^2t$,
1759: \[
1760: \Vv(G^{\la_u}(t)-G^\vr(t))\leq\frac{8}{1-\vr}\cdot\Ev(U^\vr_{Z^\vr(t)^+})+
1761: \frac{8(u+1)}{(1-\vr)^2}
1762: \]
1763: \end{lm}
1764: \begin{proof}
1765: We start with substituting \eqref{eq:mndef} into Lemma \ref{lm:vvlr}
1766: (integer parts can  be dropped without violating the inequality):
1767: \[
1768: \Vv(G^{\la_u}(t))\leq\frac{\vr^2}{\la_u^2}\cdot\Vv(G^\vr(t))+t\cdot\left(\frac{(1-\vr)^2}{(1-\la_u)^2}-\frac{\vr^2}{\la_u^2}\right).
1769: \]
1770: Utilizing  \eqref{eq:usef},
1771: %\begin{multline*}
1772: \[
1773: \frac{(1-\vr)^2}{(1-\la_u)^2}-\frac{\vr^2}{\la_u^2}%\\
1774: %\ba
1775: %&=\frac{\left(\sqrt{(1-\vr)^2-{u}/{t}}+\vr\right)^2}{(1-\vr)^2-{u}/{t}}\cdot(1-\vr)^2-\frac{\left(\sqrt{(1-\vr)^2-{u}/{t}}+\vr\right)^2}{\vr^2}\cdot\vr^2\\
1776: %&
1777: =\left(\sqrt{(1-\vr)^2-{u}/{t}}+\vr\right)^2\cdot\frac{{u}/{t}}{(1-\vr)^2-{u}/{t}}.
1778: %\ea
1779: \]
1780: %\end{multline*}
1781: Since the expression in parentheses is not larger than 1,
1782: ${u}/{t}\leq\frac34(1-\vr)^2$, and $\vr\leq\la_u$, it follows that
1783: \[
1784: \Vv(G^{\la_u}(t))\leq\Vv(G^\vr(t))+\frac{4}{(1-\vr)^2}\cdot u.
1785: \]
1786: Then we proceed with Lemma \ref{lm:ns} and \eqref{eq:mndef}:
1787: \begin{align*}
1788: &\Vv(G^{\la_u}(t)-G^\vr(t))\leq2\Vv(G^{\la_u}(t))+2\Vv(G^{\vr}(t))\\
1789: &\qquad \leq4\Vv(G^{\vr}(t))+\frac{8}{(1-\vr)^2}\cdot u\\
1790: &\qquad =\frac{8}{1-\vr}\cdot\Ev(U^\vr_{Z^\vr(t)^+})+4\frac{\fl{\vr^2t}}{\vr^2}
1791: -4\frac{\fl{(1-\vr)^2t}}{(1-\vr)^2}
1792: +\frac{8}{(1-\vr)^2}\cdot u\\
1793: &\qquad\leq\frac{8}{1-\vr}\cdot\Ev(U^\vr_{Z^\vr(t)^+})+\frac{8(u+1)}{(1-\vr)^2}
1794:  \qedhere
1795: \end{align*}
1796: \end{proof}
1797: 
1798: \begin{lm}\label{lm:vu}
1799: With the application of the coupling \eqref{eq:cou}, for any $0\leq u\leq\frac34(1-\vr)^2t$  %% (see \eqref{eq:mndef}),
1800: we have
1801: \[
1802: \Vv(U_u^{\la_u}-U_u^\vr)\leq u\cdot\frac{\vr^2}{(1-\vr)^2}.
1803: \]
1804: \end{lm}
1805: \begin{proof}
1806: By that coupling,
1807: \[
1808: \Vv[U_u^{\la_u}-U_u^\vr]=\Vv\left[\left(\frac{1-\vr}{1-\la_u}-1\right)U_u^\vr\right]=u\cdot\left(\frac{1-\vr}{1-\la_u}-1\right)^2\cdot\frac{1}{(1-\vr)^2},
1809: \]
1810: as $U^\vr_u$ is the sum of $u$ many independent $\Expd$$(1-\vr)$ weights.
1811: Write
1812: \[
1813: \ba
1814: \left(\frac{1-\vr}{1-\la_u}-1\right)\cdot\frac{1}{(1-\vr)}&=\frac{\sqrt{(1-\vr)^2-{u}/{t}}+\vr}{\sqrt{(1-\vr)^2-{u}/{t}}}-\frac{1}{(1-\vr)}\\
1815: &\leq\frac{\frac12(1+\vr)}{\frac12(1-\vr)}-\frac{1}{(1-\vr)}=\frac{\vr}{1-\vr}.
1816: \ea
1817: \]
1818: \end{proof}
1819: 
1820: \noindent
1821: After these preparations, we continue the main argument from \eqref{eq:pzu}.
1822: \begin{lm}\label{lm:zbou}
1823: There exists a constant $C_1=C_1(\vr)$ such that
1824: for any $u\geq\linebreak[4]8\vr^{-2}(1-\vr)^2$ and $t>0$,
1825: \[
1826: \Pv\{Z^\vr(t)>u\}\leq C_1\Bigl( \frac{t^2}{u^4}\cdot\Ev(U^\vr_{Z^\vr(t)^+})
1827: +\frac{t^2}{u^3}\Bigr).
1828: %\Pv\{Z^\vr(t)>u\}\les \frac{(1-\vr)^5}{\vr^2}\cdot\frac{t^2}{u^4}\cdot\Ev(U^\vr_{Z^\vr(t)^+})+\frac{(1+\vr^2)(1-\vr)^4}{\vr^2}\cdot\frac{t^2}{u^3}
1829: \]
1830: %where $\les$ means that there is a factor of the right-hand side, not depending on any of the variables, with which the inequality holds.
1831: \end{lm}
1832: \begin{proof}
1833: If $8\vr^{-2}(1-\vr)^2\leq u\leq(1-\vr)^2t$, then continuing
1834: from \eqref{eq:pzu} and taking Lemma \ref{lm:middle} into account, we write
1835: \[
1836: \ba
1837: \Pv\{Z^\vr(t)>u\}
1838: %&\leq\Pv\biggl\{U^{\la_u}_u-U^\vr_u\leq\Ev(U^{\la_u}_u-U^\vr_u)-\frac{\vr}{16(1-\vr)^3}\cdot\frac{u^2}{t}\biggr\}\\
1839: %&\quad+\Pv\biggl\{G^{\la_u}(t)-G^\vr(t)\geq\Ev(U^{\la_u}_u-U^\vr_u)-
1840: %\frac{\vr}{16(1-\vr)^3}\cdot\frac{u^2}{t}\biggr\}\\
1841: &\leq\Pv\biggl\{U^{\la_u}_u-U^\vr_u\leq\Ev(U^{\la_u}_u-U^\vr_u)-
1842: \frac{\vr}{16(1-\vr)^3}\cdot\frac{u^2}{t}\biggr\}\\
1843: &\ +\Pv\biggl\{G^{\la_u}(t)-G^\vr(t)\geq\Ev(G^{\la_u}(t)-G^\vr(t))+
1844: \frac{\vr}{16(1-\vr)^3}\cdot\frac{u^2}{t}\biggr\}\\
1845: &\leq\Vv(U^{\la_u}_u-U^\vr_u)\cdot\frac{16^2(1-\vr)^6}{\vr^2}\cdot\frac{t^2}{u^4}\\
1846: &\ +\Vv(G^{\la_u}(t)-G^\vr(t))\cdot\frac{16^2(1-\vr)^6}{\vr^2}\cdot\frac{t^2}{u^4}
1847: \ea
1848: \]
1849: by Chebyshev's inequality.  If
1850: $8\vr^{-2}(1-\vr)^2\leq u\leq\frac34(1-\vr)^2t$,
1851: use Lemmas \ref{lm:vu} and \ref{lm:vg} to conclude
1852: \begin{align*}
1853: \Pv\{Z^\vr(t)>u\}&\leq 16^2(1-\vr)^4\cdot\frac{t^2}{u^3}
1854: +8\cdot 16^2\cdot\frac{(1-\vr)^5}{\vr^2}\cdot\frac{t^2}{u^4}
1855: \cdot\Ev(U^\vr_{Z^\vr(t)^+})\\
1856: &+8\cdot 16^2\cdot\frac{(1-\vr)^4}{\vr^2}\cdot\frac{t^2(u+1)}{u^4}.
1857: \end{align*}
1858: When $\frac34(1-\vr)^2t<u\leq(1-\vr)^2t$, the previous display
1859:  works for $\frac34u$. Hence by
1860: \[
1861: \Pv\{Z^\vr(t)>u\}\leq\Pv\{Z^\vr(t)> 3u/4\},
1862: \]
1863: the statement still holds, modified by a  factor of a  power of $4/3$.
1864: 
1865: Finally, the probability is trivially zero if $u>(1-\vr)^2t$.
1866: \end{proof}
1867: 
1868: \hop
1869: Fix a number $0<\al<1$, and define
1870: \be
1871: y=\frac{u}{\al(1-\vr)}.\label{eq:uy}
1872: \ee
1873: 
1874: \begin{lm}\label{lm:ld}
1875: We have the following large deviations estimate:
1876: \[
1877: \Pv\{U^\vr_u>y\}\leq\e{-(1-\vr)(1-\sqrt\al)^2y}.
1878: \]
1879: \end{lm}
1880: \begin{proof}
1881: We use the fact that $U^\vr_u=\sum\limits_{i=1}^u\om_{i0}$, where the $\om$'s are iid.\ $\Expd$$(1-\vr)$ variables. Fix $s$ with $1-\vr>s>0$. By the Markov inequality, we write
1882: \[
1883: \ba
1884: \Pv\{U^\vr_u>y\}&=\Pv\{\e{sU^\vr_u}>\e{sy}\}\leq\e{-sy}\Ev(\e{sU^\vr_u})
1885: =\e{-sy}\cdot\Bigl(\frac{1-\vr}{1-\vr-s}\Bigr)^u\\
1886: &\leq\exp\Bigl(-sy+u\cdot\frac{s}{1-\vr-s}\Bigr).
1887: \ea
1888: \]
1889: Substituting $u=\al(1-\vr)y$, the choice $s=(1-\vr)(1-\sqrt\al)$ minimizes the exponent, and yields the result.
1890: \end{proof}
1891: 
1892: \begin{lm}
1893: There exist finite
1894: positive constants  $C_2=C_2(\al, \vr)$ and  $C_3=C_3(\al, \vr)$
1895:  such that, for all
1896: \[
1897: r\geq\frac{8(1-\vr)}{\al\vr^2 \Ev(U^\vr_{Z^\vr(t)^+})}\,,
1898: \]
1899: we have the bound
1900: \begin{multline*}
1901: \Pv\{U^\vr_{Z^\vr(t)^+}>r\Ev(U^\vr_{Z^\vr(t)^+})\}\\
1902: \leq \frac{C_2t^2}{[\Ev(U^\vr_{Z^\vr(t)^+})]^3}\cdot
1903: \left(\frac{1}{r^3}+\frac{1}{r^4}\right)
1904: +\exp\{-C_3 r\Ev(U^\vr_{Z^\vr(t)^+})\}.
1905: %\Pv\{U^\vr_{Z^\vr(t)^+}>c\Ev(U^\vr_{Z^\vr(t)^+})\}\\
1906: %\les\frac{t^2}{[\Ev(U^\vr_{Z^\vr(t)^+})]^3}\cdot\frac{1-\vr}{\al^3\vr^2}\cdot\left(\frac{1+\vr^2}{c^3}+\frac{1}{\al c^4}\right)+\e{-c(1-\vr)(1-\sqrt\al)^2\Ev(U^\vr_{Z^\vr(t)^+})}.
1907: \end{multline*}
1908: \end{lm}
1909: \begin{proof}
1910: By  \eqref{eq:uy} and  Lemmas
1911: \ref{lm:zbou} and \ref{lm:ld},  for
1912: any $ y\geq 8\al^{-1}\vr^{-2}(1-\vr)$, and with an appropriately
1913: defined new constant,
1914: \begin{align*}
1915: \Pv\{U^\vr_{Z^\vr(t)^+}>y\}&\leq\Pv\{Z^\vr(t)^+>u\}+\Pv\{U^\vr_u>y\}\\
1916: &\leq
1917: C_2 \Bigl(\,
1918: \frac{t^2}{y^4}\cdot\Ev(U^\vr_{Z^\vr(t)^+})+\frac{t^2}{y^3}\Bigr)
1919: +\e{-(1-\vr)(1-\sqrt\al)^2y}.
1920: \end{align*}
1921: Choose $y=r\Ev(U^\vr_{Z^\vr(t)^+})$.
1922: \end{proof}
1923: 
1924: \begin{tm}
1925: \[
1926: \limsup_{t\to\infty}\frac{\Ev(U^\vr_{Z^\vr(t)^+})}{t^{2/3}}<\infty,
1927: \quad\text{and}\quad\limsup_{t\to\infty}\frac{\Vv(G^\vr(t))}{t^{2/3}}<\infty.
1928: \]
1929: \label{tm:EUbd}\end{tm}
1930: \begin{proof}
1931: The first inequality implies the second one by Lemma \ref{lm:ns} and \eqref{eq:mndef}. To prove the first one, suppose that there exists a sequence $t_k\nearrow\infty$ such that
1932: \[
1933: \lim_{k\to\infty}\frac{\Ev(U^\vr_{Z^\vr(t_k)^+})}{t_k^{2/3}}=\infty.
1934: \]
1935: Then $\Ev(U^\vr_{Z^\vr(t_k)^+})>t_k^{2/3}$ for all large $k$'s,
1936: and consequently
1937: by the above lemma
1938: %\begin{multline*}
1939: \[
1940: \Pv\{U^\vr_{Z^\vr(t_k)^+}>r\Ev(U^\vr_{Z^\vr(t_k)^+})\}\\
1941: \leq C_2 \left(\frac{1}{r^3}+\frac{1}{r^4}\right)\frac{t_k^{2}}{[\Ev(U^\vr_{Z^\vr(t_k)^+})]^3}
1942: +\exp(-C_3 r t_k^{2/3})
1943: \]
1944: %\end{multline*}
1945: for all $r\geq C_4 t_k^{-2/3}$.
1946: This shows by dominated convergence that
1947: \[
1948: \int_0^\infty\Pv\{U^\vr_{Z^\vr(t_k)^+}>r
1949: \Ev(U^\vr_{Z^\vr(t_k)^+})\}\di r\underset{k\to\infty}{\longrightarrow}0,
1950: \]
1951: which leads to the contradiction
1952: \[
1953: 1 = \Ev\left(\frac{U^\vr_{Z^\vr(t_k)^+}}{\Ev(U^\vr_{Z^\vr(t_k)^+})}\right)\underset{k\to\infty}{\longrightarrow}0.
1954: \]
1955: \end{proof}
1956: 
1957: Combining Lemma \ref{lm:zbou} and Theorem \ref{tm:EUbd} gives
1958: a tail bound on $Z$:
1959: 
1960: \begin{cor} Given any $t_0>0$
1961: there exists a finite constant  $C_4=C_4(t_0, \vr)$
1962: such that, for all $a>0$ and $t\geq t_0$,
1963: \[
1964: \Pv\{ Z^\vr(t)\geq at^{2/3}\} \leq C_4 a^{-3}.
1965: \]
1966: \label{cor:Z-ub}
1967: \end{cor}
1968: 
1969: \section{Lower bound}
1970: \label{sc:lb}
1971: We abbreviate
1972:  $(m,n)=\bigl( \fl{(1-\vr)^2t},\fl{\vr^2t}\bigr)$
1973: throughout this section.
1974: %whenever it does not cause confusion.
1975: 
1976: 
1977: \begin{lm} Let $a,b>0$ be arbitrary positive numbers.
1978: There exist finite  constants $t_0=t_0(a,b,\vr)$
1979: and $C=C(\vr)$  such that, for all $t\geq t_0$,
1980: \[
1981: \Pv\Bigl\{ \,\sup_{1\leq z\leq \va t^{2/3}}\bigl(  U^\vr_z + A_z(t) -A_1(t)\bigr)
1982: \geq bt^{1/3} \Bigr\} \leq C\va^3 (b^{-3}+b^{-6}).
1983: \]
1984: \label{lm:lb-1}
1985: \end{lm}
1986: \begin{proof}
1987: The process
1988: $\{U^\vr_z\}$ depends on the boundary $\{\om_{i0}\}$.
1989: Pick a version\linebreak[4] $\{\om_{ij}\}_{1\leq i\leq m,1\leq j\leq n}$ of the  interior variables
1990: independent of $\{\om_{i0}\}$.
1991: If we use the reversed system
1992:  \be\{\wt\om_{ij}=\om_{m-i+1,n-j+1}\}_{1\leq i\leq m,1\leq j\leq n}
1993: \label{eq:reversal-1}
1994: \ee
1995:  to compute $A_z(m,n)$,
1996: then this coincides with $A_1(m-z+1,n)$ computed with $\{\om_{ij}\}$.
1997: Thus with this coupling (and some abuse of notation) we can
1998: replace $A_z(m,n) -A_1(m,n)$ with $A_1(m-z+1,n)-A_1(m,n)$.
1999: [Note that $A_1(m,n)$ is the same for $\om$ and $\wt\om$.]
2000: Next pick a further independent version of boundary conditions
2001: \eqref{eq:bondis}   with density $\la$. Use these and $\{\om_{ij}\}_{i,j\geq 1}$
2002: to compute the last-passage times $G^\la$, together with a
2003:  competition interface  $\vp^\la$ defined by
2004: \eqref{eq:def-vphi} and the projections $v^\la$
2005: defined by \eqref{eq:def-vn}. Then by  \eqref{eq:coup-4}, on the
2006: event $v^\la(n)\geq m$,
2007:  \[
2008: A_1(m,n)-A_1(m-z+1,n)  \geq G^\la(m,n)-G^\la(m-z+1,n).
2009: \]
2010:  Set
2011: \[
2012: V^\la_z= G^\la(m,n)
2013: -G^\la(m-z,n),
2014: \]
2015: a sum of $z$ i.i.d.\ Exp($1-\la$) variables. $V^\la$ is independent
2016: of $U^\vr$. Combining these steps we get the bound
2017: \be
2018: \begin{split}
2019: &\Pv\Bigl\{ \,\sup_{1\leq z\leq \va t^{2/3}}\bigl(  U^\vr_z + A_z(t) -A_1(t)\bigr)
2020: \geq bt^{1/3} \Bigr\} \\
2021: &\quad \leq
2022: \Pv\bigl\{ v^\la(\fl{\vr^2t})<\fl{(1-\vr)^2t}\bigr\}
2023:  + \Pv\Bigl\{ \,\sup_{1\leq z\leq \va t^{2/3}}(  U^\vr_z -V^\la_{z-1})
2024: \geq bt^{1/3} \Bigr\}
2025: \end{split}
2026: \label{eq:lb-temp-2}
2027: \ee
2028: 
2029:   Introduce a parameter $r>0$ whose value
2030: will be specified later, and define \be
2031: \la=\vr-rt^{-1/3}.\label{eq:def-temp-la} \ee
2032: For the second probability on the right-hand
2033: side of \eqref{eq:lb-temp-2}, define the martingale
2034: $M_z=U^\vr_z -V^\la_{z-1}-\Ev(U^\vr_z -V^\la_{z-1})$, and note that
2035: for $z\leq \va t^{2/3}$,
2036: \begin{align*}
2037: &\Ev(U^\vr_z -V^\la_{z-1})= \frac{z}{1-\vr}-\frac{z-1}{1-\la}
2038: = \frac{zrt^{-1/3}}{(1-\vr)(1-\la)}+\frac1{1-\la} \\
2039: &\leq  \frac{r\va t^{1/3}}{(1-\vr)^2} +\frac1{1-\vr}.
2040: \end{align*}
2041: As long as
2042: \be
2043:  b> r\va(1-\vr)^{-2}+t^{-1/3}(1-\vr)^{-1},
2044: \label{eq:lb-temp-2-a}
2045: \ee
2046: we get by Doob's inequality, for any $p\geq 1$,
2047: \be
2048: \begin{split}
2049: &\Pv\Bigl\{ \,\sup_{1\leq z\leq \va t^{2/3}}(  U^\vr_z -V^\la_{z-1})
2050: \geq bt^{1/3} \Bigr\}\\
2051: &\leq \Pv\Bigl\{ \,\sup_{1\leq z\leq \va t^{2/3}} M_z \geq  t^{1/3}\Bigl( b
2052:  - \frac{r\va}{(1-\vr)^2} - \frac{t^{-1/3}}{1-\vr}\,\Bigr)\,\Bigr\}\\
2053: &\leq \frac{C(p)  t^{-p/3}}{\bigl( b-r\va(1-\vr)^{-2}-t^{-1/3}(1-\vr)^{-1}\bigr)^p}
2054: \Ev\bigl[\,\lvert  M_{\fl{\va t^{2/3}}} \rvert^p\,\bigr]\\
2055: &\leq \frac{C(p,\vr)  \va^{p/2}}{\bigl( b-r\va(1-\vr)^{-2}-t^{-1/3}(1-\vr)^{-1}\bigr)^p}.
2056: \end{split}
2057: \label{eq:lb-temp-2-b}\ee
2058: Now choose $t_0=4^3b^{-3}(1-\vr)^{-3}$. Then for $t\geq t_0$ the above
2059: bound is dominated by
2060: \[
2061: \frac{C(p,\vr)  \va^{p/2}}
2062: {\Bigl( \frac{3b}4 - \frac{r\va}{(1-\vr)^2}\Bigr)^p}
2063: \]
2064: which becomes $C(p,\vr)\va^{3} b^{-6}  $
2065: once we choose
2066: \be r=\frac{b(1-\vr)^2}{4\va}
2067: \label{eq:def-temp-r-1}\ee
2068: and $p=6$,
2069: and change the constant $C(p,\vr)$.
2070: 
2071: For  the first probability on the right-hand
2072: side of \eqref{eq:lb-temp-2},
2073: introduce the time $s=(\vr/\la)^2t$. Then
2074: \[
2075: \Pv\bigl\{ v^\la(\fl{\vr^2t})<\fl{(1-\vr)^2t}\bigr\}=\Pv\bigl\{ v^\la(\fl{\la^2s})<\fl{\la^2(1-\vr)^2\vr^{-2}s}\bigr\}.
2076: \]
2077: Notice that since $\la<\vr$ here, $\fl{\la^2(1-\vr)^2\vr^{-2}s}\leq\fl{(1-\la)^2s}$ and so
2078: %the first line of
2079: by redefining \eqref{eq:char-mn} and \eqref{eq:zstar} with $s$ and $\la$, we have that the event $v^\la(\fl{\la^2s})<\fl{\la^2(1-\vr)^2\vr^{-2}s}$ is equivalent to
2080: \[
2081: \ba
2082: Z^{*\la}(s)&=[\fl{(1-\la)^2s}-v^\la(\fl{\la^2s})]^+-[\fl{\la^2s}-w^\la(\fl{(1-\la)^2s})]^+\\
2083: &=\fl{(1-\la)^2s}-v^\la(\fl{\la^2s})\\
2084: &>\fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}.
2085: \ea
2086: \]
2087: By $Z^{*\la}\overset{\text{d}}{=}Z^\la$, we conclude
2088: \be
2089: \begin{split}
2090: &\Pv\bigl\{ v^\la(\fl{\vr^2t})<\fl{(1-\vr)^2t}\bigr\} \\
2091: &\qquad
2092: =\Pv\bigl\{Z^\la(s)>\fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}\bigr\}.
2093: \end{split}
2094: \label{eq:vtoZ}
2095: \ee
2096: %\eqref{eq:Z-vphi} applies.
2097: %\begin{align*}
2098: %&\Pv\bigl\{ v^\la(\fl{\vr^2t})<\fl{(1-\vr)^2t}\bigr\} \\
2099: %&=\Pv\bigl\{ v^\la(\fl{\la^2s})<\fl{\la^2(1-\vr)^2\vr^{-2}s}\bigr\}\\
2100: %&=\Pv\bigl\{ Z^\la(s) > \fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}\bigr\}.
2101: %\end{align*}
2102: Utilizing the definitions \eqref{eq:def-temp-la} and
2103:  \eqref{eq:def-temp-r-1} of $\la$ and  $r$, one can
2104: check that by  increasing $t_0=t_0(a,b,\vr)$ if necessary, one can guarantee that for
2105:  $t\geq t_0$  there exists
2106: a constant $C=C(\vr)$ such that
2107: \[
2108: \fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}\geq Crs^{2/3}.
2109: \]
2110: Combining this  with Corollary \ref{cor:Z-ub} and definition
2111: \eqref{eq:def-temp-r-1} of $r$  we get the bound
2112: \[
2113: \begin{split}
2114: \Pv\bigl\{ v^\la(\fl{\vr^2t}) <\fl{(1-\vr)^2t}\bigr\}
2115: &\leq \Pv\bigl\{ Z^\la(s) > Crs^{2/3} \bigr\} \\
2116: &\leq Cr^{-3} \leq C (\va/b)^3.
2117: \end{split}
2118: \]
2119: 
2120: Returning to \eqref{eq:lb-temp-2} to combine
2121: all the bounds, we have
2122: \[
2123: \Pv\Bigl\{ \,\sup_{1\leq z\leq \va t^{2/3}}\bigl(  U^\vr_z + A_z(t) -A_1(t)\bigr)
2124: \geq bt^{1/3} \Bigr\}
2125:  \leq  C\Bigl( \,\frac{\va^3}{b^3}+ \frac{\va^3}{b^6}\,\Bigr).
2126: \qedhere \]
2127: \end{proof}
2128: 
2129: \begin{lm}  We have the asymptotics
2130: \[
2131: \lim_{\ve\searrow 0} \limsup_{t\to\infty}
2132: \Pv\{ 0< U^\vr_{Z^{\vr}(t)^+} \leq \ve t^{2/3} \} =0.
2133: \]
2134: \label{lm:lb-2}
2135: \end{lm}
2136: 
2137: Note that part of the event is the requirement $Z^\vr(t)>0$.
2138: 
2139: \begin{proof}
2140: The limit comes from control over the point $Z^\vr(t)$. First
2141: write
2142: \begin{align*}
2143: \Pv\{ 0< U^\vr_{Z^{\vr}(t)^+} \leq \ve t^{2/3} \}
2144: \leq \Pv\{ 0<Z^\vr(t)\leq \delta t^{2/3}\}
2145: + \Pv\{ U^\vr_{\tfl{\delta t^{2/3}}}\leq \ve t^{2/3} \}.
2146: \end{align*}
2147: Given $\delta>0$, the last probability vanishes as $t\to\infty$
2148: for any $\ve< \delta(1-\vr)^{-1}$. Thus it remains to show that
2149: the first probability on the right can be made  arbitrarily small
2150: for large $t$, by choosing a small enough $\delta$.
2151: 
2152: Let $0<\delta,b<1$.
2153: \begin{align}
2154: &\Pv\{ 0<Z^\vr(t)\leq \delta t^{2/3}\} \nn\\
2155: &\leq \Pv\Bigl\{ \,\sup_{x>\delta t^{2/3}}\bigl( U^\vr_x+A_x(t)\bigr) <
2156: \sup_{1\leq x\leq \delta t^{2/3}}\bigl( U^\vr_x+A_x(t)\bigr) \Bigr\}\nn\\
2157: &\leq \Pv\Bigl\{ \,\sup_{x>\delta t^{2/3}}
2158: \bigl( U^\vr_x+A_x(t)-A_1(t)\bigr) < bt^{1/3}\Bigr\}\label{eq:prob-temp-4}\\
2159: &\qquad +\;
2160: \Pv\Bigl\{ \,\sup_{1\leq x\leq \delta t^{2/3}}
2161: \bigl( U^\vr_x+A_x(t)-A_1(t)\bigr) >  bt^{1/3}\Bigr\}.
2162: \label{eq:prob-temp-5}
2163: \end{align}
2164: 
2165: By Lemma \ref{lm:lb-1}
2166:  the probability \eqref{eq:prob-temp-5} is bounded
2167: by $C\delta^3(b^{-3}+b^{-6})$.
2168: Bound the probability \eqref{eq:prob-temp-4} by
2169: \begin{align}
2170: &\Pv\Bigl\{ \,\sup_{\delta t^{2/3}<x\leq t^{2/3}}
2171: \bigl( U^\vr_x+A_x(t)-A_1(t)\bigr) < bt^{1/3}\Bigr\}\nn\\
2172: &\quad \leq
2173: \Pv\bigl\{ v^\la(\fl{\vr^2t})> \fl{(1-\vr)^2t}-t^{2/3}\bigr\}
2174: \label{eq:prob-temp-6}\\
2175: &\qquad  + \Pv\Bigl\{ \,\sup_{\delta t^{2/3}<x\leq t^{2/3}}
2176: (  U^\vr_x -V^\la_{x})
2177: < bt^{1/3} \Bigr\} \label{eq:prob-temp-7}
2178: \end{align}
2179: where, following the example of
2180:  the previous proof, we have introduced a new density,
2181: this time
2182: \[
2183: \la=\vr+rt^{-1/3},
2184: \]
2185: and then used the reversal trick
2186: of equation \eqref{eq:reversal-1} and Lemma \ref{lm:coup-3}
2187: to deduce
2188: \[
2189: A_x(m,n)-A_1(m,n)\geq G^\la(m-x+1,n)-G^\la(m,n)\equiv -V^\la_{x-1}
2190: \geq -V^\la_x,
2191: \]
2192: whenever $v^\la(\fl{\vr^2t})\leq \fl{(1-\vr)^2t}-t^{2/3}$. We claim that, given $\eta>0$ and parameter  $r$ from above,
2193: we can fix  $\delta, b>0$  small
2194: enough so that, for some $t_0<\infty$,  the probability in
2195: \eqref{eq:prob-temp-7} satisfies
2196: \be
2197: \Pv\Bigl\{ \,\sup_{\delta t^{2/3}<x\leq t^{2/3}}
2198: (  U^\vr_x -V^\la_{x})
2199: < bt^{1/3} \Bigr\} \leq \eta
2200: \qquad\text{for all $t\geq t_0$.}
2201: \label{eq:lb-temp-11}
2202: \ee
2203: As $t\to\infty$,
2204: \begin{align*}
2205: &t^{-1/3} \Ev(U^\vr_{\fl{yt^{2/3}}} -V^\la_{\fl{yt^{2/3}}})
2206: \longrightarrow \frac{-ry}{(1-\vr)^2}\\
2207: &\quad\quad\text{and}\quad
2208: t^{-2/3}\Vv(U^\vr_{\fl{yt^{2/3}}} -V^\la_{\fl{yt^{2/3}}})
2209: \longrightarrow \frac{2y}{(1-\vr)^2}\equiv\sigma^2(\vr)y
2210: \end{align*}
2211: uniformly over $y\in[\delta, 1]$. Since we have a sum of
2212: i.i.d's,
2213: the probability in \eqref{eq:lb-temp-11} converges, as $t\to\infty$, to
2214: \[
2215: \Pv\Bigl\{ \,\sup_{\delta \leq y\leq 1}\Bigl(
2216: \sigma(\vr) B(y)-\frac{ry}{(1-\vr)^{2}}\Bigr)
2217: \leq  b \Bigr\}
2218: \]
2219: where $B(\cdot)$ is standard Brownian motion. The random variable
2220: \[ \sup_{0\leq y\leq 1}\Bigl(
2221: \sigma(\vr) B(y)-\frac{ry}{(1-\vr)^{2}}\Bigr)
2222: \]
2223: is positive almost surely, so the above probability is less
2224: than $\eta/2$ for small $\delta$ and $b$. This implies
2225: \eqref{eq:lb-temp-11}.
2226: 
2227: The probability in \eqref{eq:prob-temp-6}
2228: is bounded by
2229: \begin{align*}
2230: &\Pv\bigl\{ v^\la(\tfl{\vr^2t})>\tfl{(1-\vr)^2t-t^{2/3}-1}\bigr\}\\
2231: &\leq  \Pv\bigl\{ v^{1-\la}(\tfl{(1-\vr)^2t-t^{2/3}-1})<\tfl{\vr^2t}\bigr\}\\
2232: &\leq \Pv\bigl\{ v^{1-\la}(\tfl{(1-\la)^2s})<\tfl{\la^2s}- qs^{2/3}\bigr\}\\
2233: &=\Pv\bigl\{ Z^{*1-\la}(s)>  qs^{2/3}\bigr\}\\
2234: &=\Pv\bigl\{ Z^{1-\la}(s)>  qs^{2/3}\bigr\}\\
2235: &\leq Cq^{-3}.
2236: \end{align*}
2237: Above we first used  \eqref{eq:v-w-ineq} and
2238:  transposition of the array $\{\om_{ij}\}$. Because this exchanges
2239: the axes, density $\la$ becomes $1-\la$.
2240:  Then we defined
2241: $s$  by
2242: \[
2243: (1-\la)^2s= (1-\vr)^2t-t^{2/3}-1
2244: \]
2245: and observed that for large enough $t$, the second inequality
2246: holds for some $q=
2247: C(\vr)((1-\vr)r-\vr )$. We used \eqref{eq:zstar} and the distributional identity of $Z$ and $Z^*$ thereafter. The last inequality is
2248: from Corollary \ref{cor:Z-ub}.
2249: 
2250: Now given $\eta>0$, choose $r$ large enough so that
2251: $Cq^{-3}<\eta$.  Given this $r$, choose $\delta, b$ small enough so
2252: that \eqref{eq:lb-temp-11} holds.  Finally, shrink $\delta$ further
2253: so that $C\delta^3(b^{-3}+b^{-6})<\eta$ (shrinking $\delta$ does not
2254: violate \eqref{eq:lb-temp-11}).
2255: To summarize, we have shown that, given $\eta>0$,
2256: if $\delta$ is small enough, then for all large  $t$
2257: \be
2258: \Pv\{1\leq  Z^\vr(t)\leq \delta t^{2/3}\}\leq 3\eta.
2259: \label{eq:b-proof}
2260: \ee
2261: This concludes the proof of the lemma.
2262: \end{proof}
2263: 
2264: Via transpositions we get the previous lemma also for the j-axis:
2265: 
2266: \begin{cor}   We have the asymptotics
2267: \[
2268: \lim_{\ve\searrow 0} \limsup_{t\to\infty}
2269: \Pv\{ 0< U^\vr_{-Z^{\vr}(t)^-} \leq \ve t^{2/3} \} =0.
2270: \]
2271: \label{cor:lb-3}
2272: \end{cor}
2273: \begin{proof} Let  $\{\om_{ij}\}$ be an initial assignment
2274: with density $\vr$. Let $\wt\om_{ij}=\om_{ji}$ be the
2275: transposed array,
2276: which is an initial assignment with density $1-\vr$.
2277: Under transposition the
2278: $\fl{(1-\vr)^2t}\times\fl{\vr^2 t}$ rectangle has become
2279: $\fl{\vr^2 t}\times\fl{(1-\vr)^2t}$, the correct characteristic
2280: dimensions for density $1-\vr$.  Since  transposition exchanges
2281: the coordinate axes, after transposition
2282: $U^\vr_{Z^\vr(t)^+}$ has become $U^{1-\vr}_{-Z^{1-\vr}(t)^-}$,
2283: and so these two random variables
2284: have the same distribution. The corollary is now
2285: a consequence of Lemma \ref{lm:lb-2} because this lemma is valid
2286: for each density $0<\vr<1$.
2287: \end{proof}
2288: 
2289: \eqref{eq:b-proof} proves part (b) of Theorem \ref{tm:Z-1}.
2290: The theorem below gives the lower bound for Theorem \ref{tm:G-var-1}
2291: and thereby completes its proof.
2292: 
2293: \begin{tm}
2294: \[
2295: \liminf_{t\to\infty}\frac{\Ev(U^\vr_{Z^\vr(t)^+})}{t^{2/3}}>0,
2296: \quad\text{and}\quad\liminf_{t\to\infty}\frac{\Vv(G^\vr(t))}{t^{2/3}}>0.
2297: \]
2298: \end{tm}
2299: \begin{proof}
2300: Suppose
2301: there exists a density $\vr$ and a sequence $t_k\to\infty$ such that
2302: $t_k^{-2/3}\Vv(G^\vr(t_k))\to 0$.  Then by Lemma \ref{lm:ns}
2303: \[
2304: \frac{\Ev(U^\vr_{Z^\vr(t_k)^+})}{t_k^{2/3}}\to 0
2305: \quad\text{and}\quad
2306: \frac{\Ev(U^\vr_{-Z^\vr(t_k)^-})}{t_k^{2/3}}\to 0.
2307: \]
2308: From this and Markov's inequality
2309: \[
2310: \Pv\{ U^\vr_{Z^\vr(t_k)^+} >\ve t_k^{2/3}\}\to 0
2311: \quad\text{and}\quad
2312: \Pv\{ U^\vr_{-Z^\vr(t_k)^-} >\ve t_k^{2/3}\}\to 0
2313: \]
2314: for every $\ve>0$. This together with Lemma \ref{lm:lb-2}
2315: and Corollary \ref{cor:lb-3} implies
2316: \[
2317: \Pv\{ U^\vr_{Z^\vr(t_k)^+} > 0\}\to 0
2318: \quad\text{and}\quad
2319: \Pv\{ U^\vr_{-Z^\vr(t_k)^-} > 0\}\to 0.
2320: \]
2321: But these statements imply that
2322: \[
2323: \Pv\{ Z^\vr(t_k) > 0\}\to 0
2324: \quad\text{and}\quad
2325: \Pv\{ Z^\vr(t_k) < 0\}\to 0,
2326: \]
2327: which is  a contradiction since these two probabilities
2328: add up to 1 for each fixed $t_k$.
2329: This proves the second claim of the theorem.
2330: 
2331: The first claim follows because it is equivalent to the second.
2332: \end{proof}
2333: 
2334: \section{Rarefaction boundary conditions}
2335: \label{sc:gen-bd}
2336: In this section we prove results on the longitudinal and transversal fluctuations of a maximal path under more general
2337: boundary conditions.
2338: %% that correspond to the case where a rarefaction fan appears,
2339: %%see e.g.\ \cite{serf}.
2340: Abbreviate as before
2341: \[ (m,n) = \bigl( \fl{(1-\vr)^2t},\fl{\vr^2t}\bigr).\]
2342: We start by studying $A_0(t)=A_0(m,n)$, the maximal path to $(m,n)$ when there are no weights on the axes. We still use the boundary conditions \eqref{eq:bondis}, so that we have coupled $A_0(t)$ and $G^\vr(t)$.
2343: We prove another version
2344: of  Lemma \ref{lm:lb-1} to make it applicable for all $t\ge 1$.
2345: \begin{lm}\label{lem:A0}
2346: Fix $0<\al <1$. There exists a constant $C=C(\al, \vr)$
2347:  such that, for each $t\ge 1$ and $b\ge C$,
2348: \[ \Pv\{G^\vr(t) - A_0(t)\geq bt^{1/3}\} \leq Cb^{-3\al/2}.\]
2349: \end{lm}
2350: \begin{proof}
2351: Note that
2352: \be
2353: \ba
2354: \Pv\{G^\vr(t) - A_0(t)\geq bt^{1/3}\}
2355: \leq & \ \Pv\{\sup_{|z|\leq at^{2/3}} U^\vr_z(t) + A_z(t) - A_0(t)\geq bt^{1/3}\}\\
2356: & + \Pv\{\sup_{|z|\leq at^{2/3}} U^\vr_z(t) + A_z(t)\neq G^\vr(t)\}.
2357: \label{eq:splitAG}
2358: \ea
2359: \ee
2360: The last term of \eqref{eq:splitAG} can easily be dealt with using
2361: Corollary \ref{cor:Z-ub}: there exists a $C=C(\vr)$ such that
2362: \be
2363: \ba
2364: &\Pv\{\sup_{|z|\leq at^{2/3}} U^\vr_z(t) + A_z(t)\neq G^\vr(t)\}
2365: \leq  \Pv\{ Z^\vr(t)\geq at^{2/3}\} \\
2366: &\qquad  + \Pv\{ Z^\vr(t)\leq -at^{2/3}\}  \leq Ca^{-3}. \label{eq:ubneq}
2367: \ea
2368: \ee
2369: For the first term of \eqref{eq:splitAG} we will use the results from the proof of Lemma \ref{lm:lb-1}. We split the range of $z$ into $[1,at^{2/3}]$ and $[-1,-at^{2/3}]$ and consider for now only the first part. Define
2370: \[ \la = \vr - rt^{-1/3}.\]
2371: We can use \eqref{eq:lb-temp-2} and \eqref{eq:lb-temp-2-b},
2372: where we choose $a=b^{\al/2}$, $p=2$,  and $r=b^{\al/2}$.
2373: Choose $C=C(\al, \vr)>0$ large enough so that
2374: for $b\ge C$
2375: \eqref{eq:lb-temp-2-a} is satisfied and the denominator
2376: of the last bound in \eqref{eq:lb-temp-2-b} is at least
2377: $b/2$. Then we can claim that, for all $b\ge C$ and $t\ge 1$,
2378: \be
2379: \ba
2380: &\Pv\{\sup_{1\leq z\leq at^{2/3}} U^\vr_z(t) + A_z(t) - A_0(t)
2381: \geq bt^{1/3}\} \\
2382: & \qquad \leq \Pv\bigl\{ v^\la(\fl{\vr^2t})<\fl{(1-\vr)^2t}\bigr\}
2383:  + Cb^{\al/2-2}.
2384: \ea
2385: \label{eq:sec7-temp-5}
2386: \ee
2387: From \eqref{eq:vtoZ} we get with $s=(\vr/\la)^2t$
2388: \[
2389: \Pv\bigl\{ v^\la(\fl{\vr^2t})<\fl{(1-\vr)^2t}\bigr\}=\Pv\bigl\{Z^\la(s)>\fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}\bigr\}.
2390: \]
2391: Now we continue differently  than in Lemma
2392: \ref{lm:lb-1} so that $t$ is not forced to be large.
2393: An elementary calculation yields
2394: \[
2395: \fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}\geq
2396: 2\,\frac{1-\vr}{\vr}\,rs^{2/3} + \frac{2\vr-1}{\vr^2}\,r^2s^{1/3}-1.
2397: \]
2398: We want to write down conditions
2399: under which  the right-hand side above is at least
2400: $\delta rs^{2/3}$ for some constant $\delta$ and all
2401: $s\ge 1$.  First increase the above constant $C=C(\al, \vr)$
2402: so that if $b=r^{2/\al}\ge C$, then
2403: \[
2404: \frac{1-\vr}{\vr}\,rs^{2/3} -1 \ge \frac{1-\vr}{2\vr}\,rs^{2/3}
2405: \qquad\text{ for all $s\ge 1$. }
2406: \]
2407: Then choose $\eta=\eta(\al,\vr)>0$ small enough such that
2408: %for all $C\le b\leq \eta t^{2/(3\al)}$
2409: whenever $b\in[C,\,\eta t^{2/(3\al)}]$
2410: %(this implies that 
2411: (in this case $r$ is small enough compared to $t^{1/3}$, but notice that the interval might as well be empty when $t$ is small),
2412: \[
2413: \frac{1-\vr}{\vr}\,rs^{2/3} \ge -\,\frac{2\vr-1}{\vr^2}\,r^2s^{1/3}.
2414: \]
2415: This last condition is vacuously true if $\vr\ge 1/2$.
2416: 
2417: Now we have for $C\le b\leq \eta t^{2/(3\al)}$ and with
2418: $\delta=(1-\vr)/(2\vr)$,
2419: \[ \fl{(1-\la)^2s}-\fl{\la^2(1-\vr)^2\vr^{-2}s}\geq \delta rs^{2/3}
2420: \qquad\text{ for all $s\ge 1$. }
2421: \]
2422: If we combine this with \eqref{eq:sec7-temp-5} and
2423:  Corollary \ref{cor:Z-ub},
2424: we can state that for all $C\le b\le \eta t^{2/(3\al)}$
2425: and $t\ge 1$,
2426: \[
2427: \Pv\{\sup_{1\leq z\leq at^{2/3}} U^\vr_z(t) + A_z(t) - A_0(t)
2428: \geq bt^{1/3}\}  \le Cb^{-3\al /2}
2429:  + Cb^{\al/2-2}.
2430: \]
2431: Same argument works (or just apply transposition)
2432:  for the values $-at^{2/3}\le z\le 1$,
2433: so this same upper bound is valid for the first probability
2434: on the right-hand side of \eqref{eq:splitAG}.
2435: 
2436: Taking \eqref{eq:ubneq} also into consideration,
2437:  at this point we have shown
2438: that whenever $C\le b\le \eta t^{2/(3\al)}$ and $t\ge 1$,
2439: \[ \Pv\{G^\vr(t) - A_0(t)\geq bt^{1/3}\} \leq \frac12C(b^{\al/2-2} +
2440: b^{-3\al /2})\leq Cb^{-3\al /2}.\]
2441: 
2442: What if $b\geq \eta t^{2/(3\al)}$? Note that
2443: \[ \Pv\{G^\vr(t) - A_0(t)\geq bt^{1/3}\} \leq \Pv\{G^\vr(t) \geq
2444: bt^{1/3}\}.\]
2445: Since $G^\vr(t)$ is the sum of two (dependent) random variables,
2446: each of which in turn is the sum of i.i.d. exponentials, and since
2447: \[
2448: \Ev(G^\vr(t)b^{-1}t^{-1/3}) \leq C(\vr, \eta )b^{\al -1}
2449: \]
2450: ($\Ev(G^\vr(t))$ is basically linear in $t$ by \eqref{eq:char-mn} and Lemma \ref{lm:NE}),
2451: we conclude that $\Pv\{G^\vr(t) - A_0(t)\geq bt^{1/3}\}$ goes to
2452: zero faster than any polynomial in $b$, if $b\geq \eta t^{2/(3\al)}$.
2453: This proves the lemma for all $b\ge C$.
2454: \end{proof}
2455: 
2456: Now we can establish that the
2457:  fluctuations of $A_0(t)$ are of order $t^{1/3}$.
2458: 
2459: \begin{cor}\label{cor:At}
2460: Fix $0<\al <1$. There exists a constant $C=C(\al, \vr)$ such
2461: that for all $a>0$ and $t\ge 1$,
2462: \[ \Pv \{ |A_0(t)-t| > at^{1/3}\} \leq Ca^{-3\al/2}.\]
2463: In particular this means that
2464: \[ \Ev(|A_0(t) - t|) = O(t^{1/3})\quad\text{and}\quad \Ev(A_0(t))=t -
2465: O(t^{1/3}).\]
2466: \end{cor}
2467: \begin{proof}
2468: Lemma \ref{lem:A0} together with Theorem \ref{tm:EUbd} implies for
2469: $a\ge C(\al,\vr)$
2470: \[
2471: \ba
2472: \Pv \{ |A_0(t)-t| > at^{1/3}\} &\leq \Pv\{G^\vr(t)-A_0(t)> at^{1/3}/2\}\\
2473: & \ \ \  + \Pv \{ |G^\vr(t)-t| > at^{1/3}/2\}\\
2474: & \leq C_1a^{-3\al/2} + C_2a^{-2}\\
2475: & \leq Ca^{-3\al/2}.
2476: \ea
2477: \]
2478: Finally, we can always increase $C$ in order to take all $0<a\leq
2479: C(\al,\vr)$ values into account.
2480: \end{proof}
2481: 
2482: 
2483: 
2484: We can also consider the fluctuations of the position of a maximal path.
2485: To this end we extend the definition of $Z(t)$, the exit point from the
2486: axes. We define $Z_l(t)$ as the $i$-coordinate of the right-most point on
2487: the horizontal line $j=l$ of the right-most maximal path to $(m,n)$ (we
2488: say right-most path, because later in this section we will consider
2489: boundary conditions that no longer necessarily have a unique longest
2490: path). We will use the notation $Z^\vr_l$ to denote the stationary
2491: situation and $Z^0_l$ to denote the situation where all the weights
2492: on the axes are zero. Note that in all cases
2493: \[ Z(t)^+ = Z_0(t).\]
2494: \begin{lm}
2495: Define $(k,l)=(\fl{(1-\vr)^2s},\fl{\vr^2s})$ for $s\leq t$. There
2496: exists a constant $C=C(\rho)$ such that for all $s\leq t$ with
2497: $t-s\geq 1$ and all $a>0$
2498: \[ \Pv\{ Z^\vr_l(t)\geq k+a(t-s)^{2/3}\} \leq Ca^{-3}.\]
2499: \end{lm}
2500: \begin{proof}
2501: There are several ways to see this, for example using time-reversal.
2502: One can also pick a new origin at $(k,l)$, and define a
2503:   last-passage model in
2504: the rectangle $[k,m]\times[l,n]$ with  boundary conditions given
2505: by $I$- and $J$-increments of the  $G$-process in the original
2506: rectangle $[0,m]\times[0,n]$.  The maximizing path in this
2507: new model connects up with the original maximizing path.
2508:   Hence in this new model
2509:  it looks as though the maximal path to
2510: $(m-k,n-l)$ exits the $i$-axis beyond the point $a(t-s)^{2/3}$, and so
2511: \[ \Pv\{ Z^\vr_l(t)\geq k+a(t-s)^{2/3}\} = \Pv\{ Z^\vr(t-s)\geq a(t-s)^{2/3}\}.\]
2512: We have ignored the integer parts here, but this can be dealt with uniformly in $a>0$. Now we can use Corollary \ref{cor:Z-ub} to conclude that
2513: \[ \Pv\{ Z^\vr_l(t)\geq k+a(t-s)^{2/3}\} \leq Ca^{-3}.\]
2514: \end{proof}
2515: 
2516: To get a similar result for $Z^0_l(t)$ we need a more convoluted
2517: argument and the conclusion is  a little weaker.
2518: 
2519: %%We would like to have a similar result for $Z^0_l(t)$, but since we have no stationarity, we cannot use time-reversal or a new origin. The following intuitive idea should work: given that $Z^0_l(t)$ is large, one would think that the probability that $Z^\vr(t)>0$ is bigger than in the case where we would have no information. If $Z^\vr(t)>0$, then $Z^\vr_l(t)\geq Z^0_l(t)$ (since the two maximal paths can coalesce, but not cross), which would give the desired bound for the probability, using the previous Lemma. However, we were not able to make this argument stick, so we had to use a more convoluted argument.
2520: \begin{lm}\label{lem:Z0sm}
2521: Define $(k,l)=(\fl{(1-\vr)^2s},\fl{\vr^2s})$ for $s\leq t$. There exists a constant $C=C(\al, \vr)$ such that for all $a>0$ and $t\geq 1$
2522: \[ \Pv\{ Z^0_l(t)> k+a\,t^{2/3}\} \leq Ca^{-3\al }.\]
2523: \end{lm}
2524: \begin{proof}
2525: The event $\{Z^0_l(t)> k+u\}$ is equivalent to the event
2526: \[ E=\left\{A_0(t) = \sup_{z\geq u}\{ A_0(k+z+1,l) + \tilde{A}_0(z-u,0)\}\right\}.\]
2527: Here, $A_0(i,j)$ is the weight of the maximal path (not using the axes) from $(0,0)$ to $(i,j)$, including the endpoint, whereas $\tilde{A}_0(i,j)$ is the weight of the maximal path from $(k+u+i,l+j)$ to $(m,n)$, including the endpoint but excluding the starting point {\em and excluding all the weights directly to the right or directly above $(k+u+i,l+j)$}. This corresponds to choosing $(k+u+i,l+j)$ as a new origin, and making sure that the axes through this origin have no weights. Note that the processes $A_0(\cdot,l)$ and $\tilde{A}_0(\cdot,0)$ are independent. The idea is to
2528: bound $A_0$ and $\tilde{A}_0$ by appropriate stationary processes
2529:  $G^\la$ and $G^{\tilde{\la}}$ to show that,
2530:  with high probability, this supremum will be too small if $u$ is too
2531: large.   We can couple the processes $G^\la$ and $G^{\tilde{\la}}$, where $\tilde{\la}>\la$, in the following way: $G^\la$ induces weights on the horizontal line $j=l$ through the increments of $G^\la$, see Lemma \ref{lm:NE}. The process $G^{\tilde{\la}}$ takes the point
2532: $(k+u,l)$ as origin and uses as boundary weights
2533: on the horizontal line $j=l$, with a slight abuse of
2534: notation, for $i\geq 1$
2535: \[ \tilde{\om}_{i0} = \frac{1-\la}{1-\tilde{\la}} I_{k+u+i+1,l}= \frac{1-\la}{1-\tilde{\la}}(G^\la(k+u+i+1,l)-G^\la(k+u+i,l)).\]
2536: These weights are independent $\Expd(1-\tilde{\la})$ random variables. The weights $\tilde{\om}_{0j}\sim
2537: \Expd(\tilde{\la})$ on the line $i=k+u$ can be chosen
2538: independently of everything else, whereas for $i,j\geq 1$
2539: \[ \tilde{\om}_{ij} = \om_{u+k+i,l+j}.\]
2540: So $G^{\tilde{\la}}(i,j)$ equals the weight of the maximal path
2541: from $(k+u,l)$ to $(k+u+i,l+j)$, using as weights on the
2542: points $(k+u+i,l)$ the $\tilde{\om}_{i0}$ (for $i\geq 1$), on the
2543: points $(k+u,l+j)$ the $\tilde{\om}_{0j}$ (for $j\geq 1$) and on
2544: the points $(k+u+i,l+j)$ the original $\om_{k+u+i,l+j}$
2545: (for $i,j\geq 1$). This construction leads to
2546: \[ A_0(i,j)\leq G^\la(i,j)\quad \text{and}\quad
2547: \tilde{A}_0(i,j)\leq G^{\tilde{\la}}(m-k-u,n-l) - G^{\tilde{\la}}(i,j).\]
2548: Also, for all $0\leq i\leq m-k-u-1$,
2549: \[ G^\la(k+u+i+1,l)-G^\la(k+u+1,l)\leq G^{\tilde{\la}}(i,0).\]
2550: Therefore, for all $0\leq i\leq m-k-u-1$,
2551: \[
2552: \ba
2553: A_0(k+u+i+1,l) + \tilde{A}_0(i,0) &\leq G^\la(k+u+i+1,l)- G^{\tilde{\la}}(i,0) + \\
2554: &\ \ \ \ G^{\tilde{\la}}(m-k-u,n-l) \\
2555: &\leq G^\la(k+u+1,l) + G^{\tilde{\la}}(m-k-u,n-l).
2556: \ea
2557: \]
2558: So we get
2559: \be
2560: \label{eq:boundPE}
2561: \Pv (E) \leq \Pv\{ A_0(t) - G^\la(k+u+1,l) -
2562: G^{\tilde{\la}}(m-k-u,n-l)\leq 0\}.
2563: \ee
2564: Here, we can still choose $\la$ and $\tilde{\la}$ as long as $0<\la <\tilde{\la}$, but it is not hard to see that for the optimal choices (in expectation) of $\la$ and $\tilde\la$ are determined by
2565: \be
2566: \frac{(1-\la)^2}{\la^2}=\frac{k+u+1}{l}\quad\text{and}\quad\frac{(1-\tilde\la)^2}{\tilde\la^2}=\frac{m-k-u}{n-l}.\label{eq:optchar}
2567: \ee
2568: With these choices we get
2569: \[ \Ev(G^\la(k+u+1,l)) = (\sqrt{k+u+1} + \sqrt{l})^2\]
2570: and
2571: \[ \Ev(G^{\tilde{\la}}(m-k-u,n-l)) = (\sqrt{m-k-u}+\sqrt{n-l})^2.\]
2572: This particular choice of $(\la, \tilde{\la})$ is valid (i.e., $\tilde{\la}>\la$) as soon as $u\geq C(\vr)$. Smaller $u$ can be dealt with by increasing $C$ in the statement of lemma. We have for $u\geq 2$
2573: \[ 
2574: \ba
2575: \Ev(G^\la(k+&u+1,l) + G^{\tilde{\la}}(m-k-u,n-l)) =  m+n+2\sqrt{l(k+u+1)}\\
2576: &  \hspace{5cm} +2\sqrt{(n-l)(m-k-u)}+1\\
2577: & \leq  ((1-\vr)^2+\vr^2)t + 1 + 2\sqrt{\vr^2s}\,\sqrt{(1-\vr)^2s+u} + \sqrt{\frac{l}{k+u}}\\
2578: & \hspace{2.5cm} + 2\sqrt{\vr^2(t-s)+1}\,\sqrt{(1-\vr)^2(t-s)-u+1}\\
2579: & \leq  ((1-\vr)^2+\vr^2)t + C(\vr) + 2\sqrt{\vr^2s}\,\sqrt{(1-\vr)^2s+u}\\
2580: & \hspace{2cm} + 2\sqrt{\vr^2(t-s)}\,\sqrt{(1-\vr)^2(t-s)-(u-1)}\\
2581: & \leq t + C(\vr) + \frac{\vr}{1-\vr}u - \frac{\vr}{1-\vr}(u-1) - \frac14\, \frac{\vr}{(1-\vr)^3}\, \frac{(u-1)^2}{t-s}\\
2582: & \leq t - C_1(\vr)\,\frac{u^2}{t} + C_2(\vr).
2583: \ea
2584: \]
2585: If $u=\fl{at^{2/3}}$, then we can choose constants $M=M(\vr)$ and $C_1=C(\vr)$ such that for all $a>M$ and $t\geq 1$,
2586: \be \label{eq:EGla}
2587: \Ev(G^\la(k+u+1,l) + G^{\tilde{\la}}(m-k-u,n-l)) \leq t - C_1 a^2t^{1/3}.
2588: \ee
2589: Smaller $a$ can be dealt with by increasing the constant $C$ in the statement of the lemma. Now note that, using (\ref{eq:boundPE}), we get 
2590: \[
2591: \ba
2592: \Pv(E) &\leq  \Pv\{ A_0(t) - t\leq G^\la(k+u+1,l) + G^{\tilde{\la}}(m-k-u,n-l) - t\}\\
2593: &\leq  \Pv( A_0(t)-t \leq -\frac12C_1a^2t^{1/3}) \\
2594: & \ \ \    + \Pv( G^\la(k+u+1,l) + G^{\tilde{\la}}(m-k-u,n-l) - t\geq -\frac12 C_1a^2t^{1/3})\\
2595: &\leq  \Pv( A_0(t)-t \leq -\frac12 C_1a^2t^{1/3}) + C_2a^{-4} \leq Ca^{-3\al}.
2596: \ea
2597: \]
2598: For the last line we used (\ref{eq:EGla}), the fact that
2599: \[ \Vv(G^\la(k+u+1,l) + G^{\tilde{\la}}(m-k-u,n-l))\leq Ct^{2/3},\]
2600: (notice that the choice \eqref{eq:optchar} places these coordinates in the $G$'s on the respective characteristics, see \eqref{eq:char-mn}), and Corollary \ref{cor:At}.
2601: 
2602: \end{proof}
2603: 
2604: We now turn to the case of a rarefaction fan introduced by \eqref{eq:rareom}.
2605: \begin{proof}[Proof of Theorem \ref{tm:rare}]
2606: The statement follows from the trivial observation that
2607: \[ A_0(t)\leq \hat{G}(t) \leq G^{\rho}(t)\]
2608: (if there is less weight, the paths get shorter), Corollary \ref{cor:At} and Theorem \ref{tm:EUbd}.
2609: \end{proof}
2610: \begin{proof}[Proof of Theorem \ref{tm:raretrans}]
2611: For the first inequality, we introduce the process $G^{\Wc=0}$, which uses the same weights as $G^\rho$, except on the $j$-axis, where all weights are 0 (so $\om^{\Wc=0}_{0j}=0$). It is not hard to see that
2612: \[ \hat{Z}_l(t) \leq Z^{\Wc=0}_l(t),\]
2613: simply because the right-most maximal path for $G^{\Wc=0}$ stays at least as long on the $i$-axis as a maximal path for $\hat{G}$, and it can coalesce with, but never cross a maximal path for $\hat{G}$. So we get
2614: \[ \Pv\{ \hat{Z}_l(t)\geq k+at^{2/3}\} \leq \Pv\{ Z^{\Wc=0}_l(t)\geq k+at^{2/3}\}.\]
2615: First we will show that with high probability, $Z^{\Wc=0}_0(t)$ is not too large. This will imply that if $Z^{\Wc=0}_l(t)$ is large, it must be because $Z^0_l(\tilde{t})$ (for an appropriately chosen $\tilde{t}$) is large, which has low probability because of Lemma \ref{lem:Z0sm}.
2616: 
2617: Note that, as in the proof of the previous lemma,
2618: \be\label{eq:ZW>u}
2619: \{ Z^{\Wc=0}_0(t)> u\} = \{ G^{\Wc=0}(t) = \sup_{z>u} (U^\vr_z + A_z(t))\}.
2620: \ee
2621: Now define a stationary process $G^\la$, with $\la > \vr$, whose origin is placed at $(u,0)$. It uses as weights on the $i$-axis
2622: \[ \om^\la_{i0} = \frac{1-\vr}{1-\la}\om_{u+i+1,0}.\]
2623: On the line $i=u$, $G^\la$ uses independent $\Expd({\la})$ weights. This construction guarantees that for $i\geq 0$
2624: \[ U^\vr_{u+i+1} - U^\vr_{u+1} \leq G^\la(i,0).\]
2625: Also, for $z>u$,
2626: \[ A_z(t) \leq G^\la(m-u,n)-G^\la(z-u-1,0).\]
2627: This implies that
2628: \[ \sup_{z>u} (U^\vr_z + A_z(t))\leq U^\vr_{u+1} + G^\la(m-u,n).\]
2629: This means that, using (\ref{eq:ZW>u}),
2630: \be\label{eq:ZWsubset}
2631: \{ Z^{\Wc=0}_0(t)> u\} \subset \{ G^{\Wc=0}(t) \leq  U^\vr_{u+1} + G^\la(m-u,n)\}.
2632: \ee
2633: Again we have that for the optimal $\la$,
2634: \[ \Ev(G^\la(m-u,n)) = (\sqrt{m-u}+\sqrt{n})^2,\]
2635: which leads to
2636: \[ 
2637: \ba
2638: \Ev(U^\vr_{u+1} &+ G^\la(m-u,n)) \leq \frac{u+1}{1-\vr} + m + n - u + 2\sqrt{n}\,\sqrt{m-u}\\
2639: & \leq (1-\vr)^2t + \vr^2 t +\frac{\vr u}{1-\vr} + C_1(\vr) + 2\sqrt{\vr^2t}\,\sqrt{(1-\vr)^2t-(u-1)}\\
2640: & \leq t + C_2(\vr) -\frac14\, \frac{\vr}{(1-\vr)^3}\,\frac{(u-1)^2}{t}.
2641: \ea
2642: \]
2643: Just as in the proof Lemma \ref{lem:Z0sm}, we see that if $u=\fl{bt^{2/3}}$, we can choose constants $M=M(\vr)$ and $C_1=C_1(\vr)$ such that for all $b>M$ and $t\geq 1$,
2644: \[ \Ev(U^\vr_{u+1} + G^\la(m-u,n)) \leq t - C_1b^2t^{1/3}.\]
2645: Note that with (\ref{eq:ZWsubset})
2646: \[
2647: \ba
2648: \Pv( Z^{\Wc=0}_0(t)> bt^{2/3}) & \leq \Pv (G^{\Wc=0}(t) -t \leq -\frac12 C_1 b^2t^{1/3})\\
2649: & \ \ \ + \Pv ( U^\vr_{u+1} + G^\la(m-u,n) -t \geq -\frac12 C_1 b^2t^{1/3}).
2650: \ea
2651: \]
2652: Now we can use the fact that
2653: \[ \Vv(U^\vr_{u+1} + G^\la(m-u,n)) = O(u+t^{2/3})\]
2654: (again, the optimal choice for $\la$ has placed the coordinates in $G$ on the characteristics w.r.t.\ $\la$),
2655: and Theorem \ref{tm:rare} to conclude that for $b>M$
2656: \[ \Pv \{Z^{\Wc=0}_0(t)> bt^{2/3}\}\leq Cb^{-3\al}.\]
2657: For $b\leq M$ we can increase $C$.
2658: 
2659: A little picture reveals that if $Z^{\Wc=0}_l(t)\geq k+at^{2/3}$ and $Z^{\Wc=0}_0(t)\leq at^{2/3}/2$, then the maximal path that does not use the weights on the axes from the point $(at^{2/3}/2,0)$ to $(m,n)$, must pass to the right of $(k+at^{2/3},l)$, an event with smaller probability than the event $\{ Z^0_l(t)\geq k + at^{2/3}/2\}$, which with Lemma \ref{lem:Z0sm} proves the first inequality.
2660: 
2661: The second inequality of the Theorem is a corollary of the first. We assume $k-at^{2/3}\ge0$, otherwise this statement is trivial. Also, we prove for $a>2\frac{(1-\vr)^2}{\vr^2}$, one can always increase $C$ if this is not the case. Fix $\wt s$ such that
2662: \[
2663: k-at^{2/3}=(1-\vr)^2\wt s,\quad\text{then}\quad k':\,=\fl{(1-\vr)^2\wt s},\qquad l':\,=\fl{\vr^2\wt s}.
2664: \]
2665: With these definitions,
2666: \[
2667: l\ge l'+\vr^2s-1-\frac{\vr^2}{(1-\vr)^2}\cdot(1-\vr)^2\wt s\ge l'+\frac{\vr^2}{(1-\vr)^2}\cdot at^{2/3}-1.
2668: \]
2669: Define also $\hat{Y}^T_{k'}$ to be the highest point of the left-most maximal path on the vertical line $i=k'$. As the left-most maximal path is North-East, we have
2670: \[
2671: \Pv\{\hat{Y}_l(t)\le k-at^{2/3}\}=\Pv\{\hat{Y}^T_{k'}\ge l\}\le\Pv\{\hat{Y}^T_{k'}\ge l'+\frac{\vr^2}{(1-\vr)^2}\cdot at^{2/3}-1\}.
2672: \]
2673: Pick $\wt a=a\vr^2/(1-\vr)^2-1>1$, then the right hand-side is bounded by $\Pv\{\hat{Y}^T_{k'}\ge l'+\wt at^{2/3}\}$. The transposed array $\wt\om_{ij}:\,=\om_{ji}$, $i,\,j\ge0$ has rarefaction fan boundary conditions w.r.t.\ the parameter $1-\vr$. Moreover, $\hat{Y}^T_{k'}$ becomes the right-most point of the right-most maximal path on the horizontal line $i=k'$ in the transpose picture. The first part of the Theorem with $1-\vr$, $\wt s\le t$ and $\wt a$ then completes the proof by $\wt a^{-3\al}<[\frac12(\wt a+1)]^{-3\al}=C'(\vr,\,\al)\cdot a^{-3\al}$.
2674: \end{proof}
2675: \bibliography{refsmarton}
2676: \bibliographystyle{plain}
2677: 
2678: \bigskip
2679: {\sc M.\ Bal\'azs, Mathematics Department, University of Wisconsin-Madison,} Van Vleck Hall, 480 Lincoln Dr, Madison WI 53706-1388, USA.\\
2680: \indent
2681: {\it E-mail address:} {\tt balazs@math.wisc.edu}
2682: 
2683: \bigskip
2684: {\sc E. Cator, Delft University of Technology, faculty EWI,} Mekelweg 4, 2628CD, Delft, The Netherlands.\\
2685: \indent
2686: {\it E-mail address:} {\tt e.a.cator@ewi.tudelft.nl}
2687: 
2688: \bigskip
2689: {\sc T.\ Sepp\"al\"ainen, Mathematics Department, University of Wis\-con\-sin-Madison,} Van Vleck Hall, 480 Lincoln Dr, Madison WI 53706-1388, USA.\\
2690: \indent
2691: {\it E-mail address:} {\tt seppalai@math.wisc.edu}
2692: 
2693: \end{document}
2694: