math0504586/s.tex
1: \date{} 
2: \title{Quantitative noise sensitivity\\and exceptional times for percolation}
3: \author{Oded Schramm \and Jeffrey E. Steif
4: \thanks{Research supported by the
5:      Swedish Natural Science Research Council
6: and the G\"{o}ran Gustafsson Foundation (KVA).}}
7: 
8: 
9: \documentclass[12pt,naturalnames]{article}
10: \usepackage{amsmath}
11: \usepackage{amssymb}
12: \usepackage{amsthm}
13: \usepackage{amsfonts}
14: \usepackage{graphicx}
15: \input labelfig.tex
16: \input epsf.tex
17: 
18: \newif\ifhyper\IfFileExists{hyperref.sty}{\hypertrue}{\hyperfalse}
19: 
20: \ifhyper\usepackage{hyperref}\fi
21: 
22: \newif\ifdraft
23: \drafttrue
24: \def\note#1{\ifdraft {\bf [#1]}\fi}
25: \long\def\comment#1{}
26: \long\def\oldVersion#1{}
27: \def\dhrule{\bigskip\hrule\smallskip\hrule\bigskip}
28: \numberwithin{equation}{section}
29: \numberwithin{figure}{section}
30: 
31: \newtheorem{theorem}{Theorem}
32: \numberwithin{theorem}{section}
33: \newtheorem{corollary}[theorem]{Corollary}
34: \newtheorem{lemma}[theorem]{Lemma}
35: \newtheorem{proposition}[theorem]{Proposition}
36: \newtheorem{conjecture}[theorem]{Conjecture}
37: \newtheorem{problem}[theorem]{Problem}
38: \newtheorem{exercise}[theorem]{Exercise}
39: \theoremstyle{remark}\newtheorem{definition}[theorem]{Definition}
40: \theoremstyle{remark}\newtheorem{remark}[theorem]{Remark}
41: \def\eref#1{(\ref{#1})}
42: \def\QED{\qed\medskip}
43: \newcommand{\Prob} {{\bf P}}
44: \newcommand{\R}{\mathbb{R}}
45: \newcommand{\C}{\mathbb{C}}
46: \newcommand{\Z}{\mathbb{Z}}
47: \newcommand{\N}{\mathbb{N}}
48: \def\diam{\mathop{\mathrm{diam}}}
49: \def\area{\mathop{\mathrm{area}}}
50: \def\dist{\mathop{\mathrm{dist}}}
51: \def\length{\mathop{\mathrm{length}}}
52: \def\ceil#1{\lceil{#1}\rceil}
53: \def\floor#1{\lfloor{#1}\rfloor}
54: \def\Im{{\rm Im}\,}
55: \def\calC{{\cal C}}
56: \def\calN{{\cal N}}
57: \def\bPsi{\mbox{\boldmath $\Psi$}} 
58: \def\Re{{\rm Re}\,}
59: \def\FC(#1,#2){{ 0 \stackrel{#2}{\leftrightarrow} #1 }}
60: \def\Ito/{It\^o}
61: \def \eps {\epsilon}
62: \def \P {{\bf P}}
63: \def\md{\mid}
64: \def\Bb#1#2{{\def\md{\bigm| }#1\bigl[#2\bigr]}}
65: \def\BB#1#2{{\def\md{\Bigm| }#1\Bigl[#2\Bigr]}}
66: \def\Bs#1#2{{\def\md{\mid}#1[#2]}}
67: \def\Pb{\Bb\P}
68: \def\Eb{\Bb\E}
69: \def\PB{\BB\P}
70: \def\EB{\BB\E}
71: \def\Ps{\Bs\P}
72: \def\Es{\Bs\E}
73: \def \prob {{\bf P}}
74: \def \expect {{\bf E}}
75: \def \p {{\partial}}
76: \def \E {{\bf E}}
77: \def\defeq{:=}
78: \def\var{\operatorname{var}}
79: \def\closure{\overline}
80: \def\ev#1{{\mathcal{#1}}}
81: \def \proof {{ \medbreak \noindent {\bf Proof.} }}
82: \def\proofof#1{{ \medbreak \noindent {\bf Proof of #1.} }}
83: \def\proofcont#1{{ \medbreak \noindent {\bf Proof of #1, continued.} }}
84: \def\CCa{C_3}
85: \def\CCb{C_1}
86: \def\CCc{C_2}
87: 
88: \def\bl{\bigl}\def\br{\bigr}\def\Bl{\Bigl}\def\Br{\Bigr}
89: \def\Vbdry{V_{\p}}
90: \def\V#1{V(#1)}
91: \def\Vss{V_{s,R}\cap V_{s',R}}
92: \def\Vb#1{V_\p(#1)}
93: \def\VD{V_D}
94: \def\sign{\mathrm{sign}}
95: 
96: \def\nn{[n]}
97: \def\cro{Q}
98: \def\croRr{f_r^R}
99: \def\mfh{\mathfrak H}
100: \def\hmfh{\hat\mfh}
101: \def\interior{\operatorname{interior}}
102: \def\II{{J}}
103: %
104: 
105: \def\noopsort#1{}
106: 
107: \hfuzz 1pt
108: %
109: \begin{document}
110: \maketitle
111: 
112:  \begin{abstract}
113: One goal of this paper is to
114:  prove that dynamical critical site percolation on the planar triangular lattice
115: has exceptional times at which percolation occurs. 
116: In doing so, new {\sl quantitative} noise sensitivity results
117: for percolation are obtained.
118: The latter is based on a novel method for controlling the
119:  ``level $k$'' Fourier coefficients via the construction of a
120: randomized 
121: algorithm which looks at random bits, outputs the value of a 
122: particular function but looks at any fixed input bit with low 
123: probability. We also obtain upper and lower bounds on the Hausdorff
124: dimension of the set of percolating times. We then study the problem
125: of exceptional times for certain
126: ``$k$-arm'' events on wedges and cones. As a corollary
127: of this analysis, we prove, among other things, that there are no times 
128: at which there are two infinite ``white'' clusters, obtain an upper bound 
129: on the Hausdorff dimension of the set of times at which there are 
130: both an infinite white cluster and an infinite black cluster
131: and prove that for dynamical critical bond percolation on the square grid
132: there are no exceptional times at which $3$ disjoint infinite clusters are
133: present.
134: \end{abstract}
135: 
136: \newpage
137: \tableofcontents
138: \newpage
139: 
140: \section {Introduction}
141: 
142: Consider bond percolation on an infinite connected locally finite graph
143: $G$, where for some $p\in[0,1]$, each edge (bond) of $G$ is, independently of 
144: all others, open with probability $p$ and closed with probability $1-p$. 
145: Write $\pi_p$ for this product measure. 
146: The main questions in percolation theory (see \cite{Grimmett})
147: deal with the possible existence of infinite connected components 
148: (clusters) in the random subgraph of $G$
149: consisting of all sites and all open edges. 
150: Write $\calC$ for the event
151: that there exists such an infinite cluster. By Kolmogorov's 0-1 law, 
152: the probability of $\calC$ is, for fixed $G$ and $p$, either 0 or 1. 
153: Since $\pi_p(\calC)$ is nondecreasing in $p$, there 
154: exists a critical probability $p_c=p_c(G)\in[0,1]$ such that
155: \[
156: \pi_p(\calC)=\left\{
157: \begin{array}{ll}
158: 0 & \mbox{for } p<p_c \\
159: 1 & \mbox{for } p>p_c.
160: \end{array} \right. 
161: \]
162: At $p=p_c$ we can have either $\pi_p(\calC)=0$ or $\pi_p(\calC)=1$, depending on $G$.
163: 
164: 
165: H\"{a}ggstr\"{o}m, Peres and Steif~\cite{HPS} initiated the study of 
166: dynamical percolation.
167: (The notion of dynamical percolation
168: was invented independently by I.~Benjamini.
169: While the present paper was motivated by \cite{HPS},
170: the question studied here had previously been asked by Benjamini, as
171: we recently became aware.)
172: In this model, with $p$ fixed, the edges of $G$ switch back 
173: and forth according to independent 2 state
174: continuous time Markov chains where closed switches to open at rate $p$ 
175: and open switches to closed at rate $1-p$. Clearly $\pi_p$ is a stationary
176: distribution for this Markov process. The general question studied in
177: \cite{HPS} was whether, when we start with distribution $\pi_p$,
178: there could exist atypical times at which the percolation 
179: structure looks markedly different than at a fixed time.
180: 
181: Write $\bPsi_p$ for the underlying probability measure
182: of this Markov process, and write $\calC_t$ for
183: the event that there is an infinite cluster of open edges 
184: at time $t$.
185: 
186: Two results in \cite{HPS} which are relevant to us are
187: 
188: \begin{proposition} \label{pr:noncrit}
189:  For any graph $G$ we have 
190: \begin{equation} \left\{ \begin{array}{ccl}
191: \bPsi_p(\, \calC_t \, \mbox{ occurs for every  } \, t \, )=1  & \mbox{ if } & p>p_c(G) 
192:    \\[1ex]
193: \bPsi_p\bigl((\neg\, \calC_t) \mbox{ occurs for every } t\bigr)=1 & \mbox{ if } & p<p_c(G) \, .
194: \end{array} \right.
195: \nonumber
196: \end{equation} 
197: \end{proposition}
198: 
199: \begin{theorem} \label{th:Zd}
200: For $d\ge 19,$ the integer lattice $\Z^d$ satisfies
201: $$
202: \bPsi_{p_c}\bl((\neg \calC_t) \mbox{ occurs for every } t\br)=1.
203: $$ 
204: \end{theorem}
205: 
206: One important aspect of the proof of the latter result is that it uses the fact,
207: proved in \cite{HS},
208: that for $d\ge 19$, 
209: \begin{equation}\label{e.z19}
210: \pi_p(0\text{ is in an infinite open cluster})
211: =O(|p-p_c|).
212: \end{equation}
213: It is proved  in \cite{KZ} that~\eref{e.z19} does not hold
214: for $d=2$.
215: Therefore, the question of whether Theorem 
216: \ref{th:Zd} is true for $d=2$ becomes interesting. At this point, we mention
217: that site percolation is the analogous model where the vertices (rather than the edges)
218: are open or closed independently each with probability $p$ and
219: dynamical percolation is defined in a completely analogous manner.
220: Our main result says that Theorem~\ref{th:Zd} does not hold
221: for site percolation on  the planar triangular grid.
222: The triangular grid is the graph whose vertex set
223: is the subset of $\C=\R^2$ consisting of the points
224: $$
225: \Z+\exp(2\,\pi\,i/3)\,\Z=
226: \bigl\{(k+\ell/2,\sqrt 3\, \ell/2):k,\ell\in\Z\bigr\}
227: $$
228: and two such points have an edge between them if and only if their distance is 1.
229: Explicitly stated, our main result is
230: 
231: 
232: \begin{theorem} \label{th:except}
233: Almost surely, the set of times
234: $t\in[0,1]$ such that dynamical critical site percolation
235: on the triangular lattice has an infinite open cluster is nonempty.
236: \end{theorem}
237: 
238: There are no other transitive graphs for which it is known that
239: dynamical critical percolation has such exceptional times.
240: (In~\cite{HPS}, it was argued that the event discussed in Theorem~\ref{th:except}
241: is measurable. A similar comment applies to our other results below.
242: Thus, measurability issues
243: will not concern us here.)
244: 
245: We are convinced that Theorem~\ref{th:except}
246: is true for bond percolation on the square lattice. However, our proof uses 
247: the existence and exact values of certain so-called 
248: {\sl critical exponents}, which
249: are only known to hold for site percolation on the triangular lattice.
250: These are believed to be the same for 
251: bond percolation on the square lattice, but even
252: their existence has not yet been established in that case. 
253: However, the methods of this paper seem to come quite close
254: to a proof for the square grid as well: it seems that there are several
255: ways in which this can perhaps be achieved without determining these
256: critical exponents. These issues will be further discussed in Section~\ref{sec:square}.
257: 
258: It is interesting to note that by~\cite[Corollary 4.2]{HPS}, a.s.\ at every
259: time $t$ the set of vertices that are contained in some infinite cluster
260: has zero density.
261: \bigskip
262: 
263: On a heuristic level, for Theorem~\ref{th:except} to hold,
264: it is necessary that
265: the configuration ``changes fast'' in order to have ``many chances''
266: to percolate  so that we will in fact have a percolating time. 
267: Mathematically, ``changing fast'' can be interpreted as having small
268: correlations over short time intervals, which then suggests the
269: use of the second moment method which we indeed will use. In other words,
270: one needs to know that the configuration at a given time tells us almost
271: nothing about how it will look a short time later. The notion of
272: ``noise sensitivity'' introduced in \cite{BKS} 
273: is the relevant tool which
274: describes this phenomenon. We now briefly explain this.
275: 
276: Given an integer $m$, a subset $A$ of $\{0,1\}^m$ and an $\eps>0$, define
277: $$
278: N(A,\eps):=
279: \var\Bigl[\Pb{(Y_1,\ldots,Y_{m})\in A\md X_1,\ldots,X_{m}}\Bigr]
280: $$
281: where 
282: $\{X_i\}_{1\le i\le m}$ are i.i.d.\ with
283: $\Pb{X_i=1}=1/2=\Pb{X_i=0}$ and conditional on the $\{X_i\}$'s,
284: $\{Y_i\}_{1\le i\le m}$ are independent with
285: $Y_i=X_i$ with probability $1-\epsilon$ and
286: $Y_i=1-X_i$ with probability $\epsilon$. 
287: 
288: \begin{definition} \label{def:noise}
289: Let $\{n_m\}_{m\ge 1}$ be an increasing sequence in $\N$ going to $\infty$ and 
290: let $A_m$ be a subset of $\{0,1\}^{n_m}$ for each $m$. We say that
291: the sequence $\{A_m\}_{m\ge 1}$ is {\bf noise sensitive} if for every $\epsilon >0$,
292: \begin{equation}\label{e.noisedef}
293: \lim_{m\to\infty} N(A_m,\eps)=0.
294: \end{equation}
295: \end{definition}
296: 
297: This says that for large $m$ knowing the values of
298: $X_1,\ldots,X_{n_m}$ gives us almost no information concerning whether
299: $(Y_1,\ldots,Y_{n_m})\in A_m$. This is not the exact definition of noise sensitivity
300: given in \cite{BKS} 
301: but is easily shown to be equivalent; see page 14 in that paper. It is also
302: shown in \cite{BKS} that if (\ref{e.noisedef}) holds for some 
303: $\epsilon \in(0, 1/2)$, then it holds for all such $\epsilon$
304: and in addition that $N(A,\eps)$ is decreasing in $\epsilon$
305: on $[0,1/2]$.
306: 
307: Let $n_m$ be the number of edges in an $(m+1)\times m$ box in $\Z^2$ and let $A_m$ be the event
308: of a left to right crossing in such a box.
309: By duality, $\Pb{A_m}=1/2$ for every $m$
310: (see~\cite{Grimmett}).
311: %(see \cite{Grimmett}, page 294). 
312: In \cite{BKS}, the following result is proved.
313: 
314: \begin{theorem} \label{th:crossing}
315: The sequence $\{A_m\}_{m\ge 1}$ is noise sensitive.
316: \end{theorem}
317: 
318: A by-product of the tools needed to prove
319:  Theorem \ref{th:except} will imply the following more quantitative
320: version of Theorem~\ref{th:crossing}, which was conjectured in~\cite{BKS}.
321: 
322: \begin{theorem} \label{th:crossingquantsquare}
323: There exists $\gamma >0$ so that
324: $$
325: \lim_{m\to\infty} N(A_m,m^{-\gamma})=0.
326: $$
327: \end{theorem}
328: 
329: 
330: We have the same result for the triangular lattice but with a better $\gamma$, since
331: critical exponents are known in this case.
332: 
333: \begin{theorem} \label{th:crossingquant} For critical
334: site percolation on the triangular lattice,
335: let $A_m'$ be the event of the existence of a left-right crossing
336: in a domain $D$ approximating a square of sidelength $m$.
337: Then for all $\gamma < 1/8$,
338: $$
339: \lim_{m\to\infty} N(A_m',m^{-\gamma})=0.
340: $$
341: \end{theorem}
342: 
343: In proving our quantitative noise sensitivity results
344: (Theorems \ref{th:crossingquantsquare} and \ref{th:crossingquant}
345: as well as those later on necessary for obtaining Theorem \ref{th:except}),
346: one of two key steps will be Theorem \ref{t.noise}, which gives 
347: estimates of certain quantities involving
348: Fourier coefficients of a function based on the properties of an algorithm
349: calculating the function;
350: the other key step will be the construction of an appropriate algorithm.
351:  Precise definitions of undefined terms 
352: will be given in Section \ref{sec:dilute},
353: where the connection with noise sensitivity will also be recalled. 
354: 
355: \begin{theorem}\label{t.noise}
356: Let $n\in\N$ and set $\Omega=\Omega_n:=\{0,1\}^n$.
357: Let $f:\Omega\to\R$ be a function.
358: Suppose that there is a randomized algorithm $A$ for
359: determining the value of $f$ which examines some of the input bits of
360: $f$ one by one, where the choice of the next bit examined may depend
361: on the bits examined so far. Let $\II\subseteq \nn:=\{1,2,\dots,n\}$
362: be the (random) set of bits examined by the algorithm.
363: Set $\delta=\delta_A:=\sup\bl\{\Ps{i\in \II}:i\in\nn\br\}$.
364: Then, for every $k=1,2,\dots$,
365: the Fourier coefficients of $f$ satisfy
366: \begin{equation}\label{e.noise}
367: \sum_{S\subseteq\nn,\,|S|=k}\hat f(S)^2 \le \delta\,k\,\|f\|^2,
368: \end{equation}
369: where $\|f\|$ denotes the $L^2$ norm of $f$ with respect to the
370: uniform probability measure on $\Omega$.
371: \end{theorem}
372: 
373: This result might have some applications to 
374: theoretical computer science. We will call $\delta_A$
375: the {\bf revealment} of the algorithm $A$. 
376: The restriction of $x$ to $J$ (the set of bits examined by the algorithm)
377: is a {\bf witness} for the function $f$, in the sense that it determines
378: $f(x)$. As explained in Section \ref{ss.noisetheorem}, Theorem~\ref{t.noise} extends to some
379: other types of witnesses.
380: 
381: In the case $k=1$, the inequality~\eref{e.noise} cannot be improved
382: by more than a factor of $O(1/\log n)$:
383: there is an example showing this with $\delta\le n^{-1/3}\,\log(n)$,
384: which appears in~\cite[\S 4]{BSW}.
385: The paper~\cite{BSW} investigates how small the revealment can be for
386: a balanced boolean function on $\{0,1\}^n$. When the function is
387: monotone, it is shown that the revealment cannot be much smaller than
388: $n^{-1/3}$ and in general it cannot be much smaller than
389: $n^{-1/2}$. Examples are given there which come within logarithmic
390: factors of meeting these bounds.
391: 
392: We don't know if~\eref{e.noise} is close to being optimal for $k\gg 1$.
393: One is tempted to speculate that the inequality can be improved
394: to $\sum_{|S|\le k}\hat f(S)^2\le O(1)\,k\,\delta\,\|f\|^2$.
395: We do not know any counterexample to this inequality.
396: However, the AND function $f(x)=\prod_{j=1}^n x_j$
397: gives an example where
398: $$
399: O(1)\,\sum_{|S|= k}\hat f(S)^2 \ge\sqrt k\, \delta\,\|f\|^2
400: $$
401: for $k$ satisfying $|k-n/2|=O(n^{1/2})$.
402: (It is easy to check that the best revealment 
403: possible for this $f$ is exactly $(2-2^{1-n})/n$.)
404: 
405: \bigskip
406: 
407: 
408: Once Theorem \ref{th:except} is established, it is natural to ask: how
409: large is the set of ``exceptional'' times at which percolation occurs?
410: In this direction, we have the following result.
411: 
412: \begin{theorem}\label{th.hd}
413: The Hausdorff dimension of the set of times 
414: at which dynamical critical site percolation
415: on the triangular lattice has an infinite cluster is 
416: an almost sure constant which lies in $[\frac{1}{6},\frac{31}{36}]$.
417: \end{theorem}
418: 
419: We conjecture that $\frac{31}{36}$ is the correct answer.
420: In a different direction, once we know that there are exceptional times at which 
421: percolation occurs, it is natural to ask how many clusters can exist at these
422: exceptional times. The following provides the answer. 
423: 
424: \begin{theorem} \label{th.no2clusters}
425: On the triangular lattice, a.s.\ there are no times at which 
426: dynamical critical site percolation has 2 or more infinite open clusters.
427: \end{theorem}
428: 
429: For the square grid, we can only prove
430: \begin{theorem} \label{th.no3clustersZ2}
431: On $\Z^2$, a.s.\ there are no times at which 
432: dynamical critical bond percolation has 3 or more infinite open clusters.
433: \end{theorem}
434: 
435: In some of the figures, we will represent open sites by white hexagons on the
436: dual grid, and
437: closed sites by black hexagons. Thus, percolation clusters
438: correspond to connected components of the union of the white hexagons.
439: These will also be called white clusters. Likewise, we may also consider
440: black clusters, which are connected components of black hexagons.
441: 
442: Asking whether 2 infinite white clusters can coexist
443: at some time is very different from asking whether 2 infinite clusters of 
444: {\sl different} colors can coexist at some time. We conjecture that there are 
445: in fact exceptional times 
446: at which there is both a white and a black infinite cluster and that the 
447: Hausdorff dimension of such times is $2/3$. We can however prove the following.
448: 
449: \begin{theorem} \label{th.differentclusters}
450: On the triangular lattice, a.s.\ the
451: Hausdorff dimension of the set of times at which there is both an infinite white
452: cluster and an infinite black
453: cluster is at most $2/3$.
454: \end{theorem}
455: 
456: We also have the following two results concerning the upper half plane.
457: 
458: \begin{theorem} \label{th.halfplane}
459: On the triangular lattice intersected with the upper half plane, a.s.\ 
460: the Hausdorff dimension of the set of times at which there is an
461: infinite cluster is at most $5/9$.
462: \end{theorem}
463: 
464: \begin{theorem} \label{th.halfplane2}
465: On the triangular lattice intersected with the upper half plane, a.s.\ the
466: set of times at which there is both an infinite white cluster 
467: and an infinite black cluster
468: is empty.
469: \end{theorem}
470: 
471: Theorems~\ref{th.no2clusters}, \ref{th.differentclusters},
472: \ref{th.halfplane} and \ref{th.halfplane2}
473: will follow immediately from generalizations presented
474: in the last part of the paper, which are concerned with studying dynamical percolation
475: on two other 2 dimensional objects, namely wedges and cones.
476: For every $\theta \in (0,\infty)$, we let $W_\theta$ denote the 
477: wedge of angle $\theta$ and $C_\theta$ denote the cone of angle $\theta$. 
478: For $C_\theta$, we will require that $\theta$ is a multiple of $\pi/3$.
479: The precise definitions of these will be given in Section \ref{sec:background}.
480: First, we mention that for all $\theta$, the critical value for site percolation
481: on $W_\theta$ and on $C_\theta$ is $1/2$, as for site percolation
482: on the triangular grid and bond percolation on $\Z^2$. 
483: 
484: The following results provide upper and lower bounds on the critical 
485: angle for which
486: there are exceptional times for certain $k$-arm type events as well as provide
487: estimates for the Hausdorff dimension of the set of exception times for a given
488: angle. In these results, if an upper bound on the Hausdorff dimension
489: is negative, this means that the set in question is empty. 
490: 
491: We will only do the case where the arms are alternating in color (and hence
492: for the case of cones, there will be one or an even number of arms).
493: We do this partially because
494: it is easier than the general case and because it is all that is needed
495: in order to make statements concerning the number of infinite clusters.
496: 
497: By a $k$-arm event, we mean an event of the form ``there are $k$ disjoint 
498: infinite paths having a specified color sequence''; 
499: for a wedge, the color sequence is well-defined while
500: for a cone, it is well-defined up to cyclic permutations.
501: 
502: 
503: \begin{theorem} \label{th:wedges}
504: Fix the wedge $W_\theta$ and for integer $k\ge 1$, let 
505: $A_{W_\theta}^k$ be the event that
506: there are $k$ infinite disjoint paths
507: in $W_\theta$ whose colors alternate. Then a.s.\ the Hausdorff 
508: dimension, $H_{W_\theta}^k$, of the set of exceptional times at which 
509: $A_{W_\theta}^k$ occurs satisfies
510: $$
511: 1-\frac{4k(k+1)\pi}{3\theta} \le  
512: H_{W_\theta}^k
513: \le
514: 1-\frac{2k(k+1)\pi}{9\theta}.
515: $$
516: In particular, for any $k\ge 1$, there are exceptional times
517: for the event $A_{W_\theta}^k$
518: for $\theta > \frac{4k(k+1)\pi}{3}$
519: and there are no exceptional times
520: for $\theta < \frac{2k(k+1)\pi}{9}$.
521: \end{theorem}
522: 
523: 
524: \begin{theorem} \label{th:cones}
525: Fix the cone $C_\theta$ with $\theta$ a multiple of $\pi/3$
526: and let, for $k= 1$ or $k>1$ even, 
527: $A_{C_\theta}^k$ be the event that
528: there are $k$ infinite disjoint paths in $C_\theta$
529: whose colors alternate (if $k>1$). 
530: Then a.s.\ the Hausdorff 
531: dimension, $H_{C_\theta}^k$, of the set of exceptional times at 
532: which $A_{C_\theta}^k$ occurs satisfies
533: $$
534: 1-\frac{5\pi}{3\theta} \le  
535: H_{C_\theta}^1
536: \le
537: 1-\frac{5\pi}{18\theta}
538: $$
539: and for $k\ge 2$
540: $$
541: 1-\frac{4(k^2-1)\pi}{3\theta} \le  
542: H_{C_\theta}^k
543: \le
544: 1-\frac{2(k^2-1)\pi}{9\theta}.
545: $$
546: In particular, for $k=1$, there are exceptional times
547: for the event $A_{C_\theta}^1$ for $\theta > \frac{5\pi}{3}$
548: and there are no exceptional times for $\theta < \frac{5\pi}{18}$,
549: while for $k\ge 2$, there are exceptional times
550: for the event $A_{C_\theta}^k$ for $\theta > \frac{4(k^2-1)\pi}{3}$
551: and there are no exceptional times for $\theta < \frac{2(k^2-1)\pi}{9}$.
552: \end{theorem}
553: 
554: Theorem \ref{th:cones} is presumably true for other values of $\theta$ 
555: provided that a proper definition of $C_\theta$ would be given.
556: 
557: \noindent
558: Remark:
559: One should note that the upper bounds on the Hausdorff
560: dimension given in Theorems~\ref{th.hd},\ref{th.differentclusters},\ref{th.halfplane},
561: \ref{th:wedges} and \ref{th:cones} are all of the form $1-(4/3)\xi$ where $\xi$ is the 
562: critical exponent for the given event.
563: 
564: \medskip
565: There is an abstract theory of L\'evy processes
566: on groups~\cite{Haw84,Eva89}, which gives a
567: criterion for a L\'evy process (such as $\omega_t$) to hit a set $A$
568: (such as the set of configurations which contain an infinite component). 
569: Basically, to show that $A$ is hit, one needs to prove that there exists a
570: probability measure $\mu$ on $A$ which has $\|\mu\|_*<\infty$
571: for an appropriate Hilbert norm $\|\cdot\|_*$, based on the Fourier
572: transform. It seems that we could use this framework in the present
573: paper, but that would not essentially simplify the core issues we deal with.
574: Moreover, it seems that our hands-on approach facilitates some generalizations,
575: which the L\'evy process theory does not cover, which brings us to
576: our next remark.
577: \medskip
578: 
579: The fact that the time between flips has an exponential
580: distribution is not really essential here,
581: and the results apply in greater generality.
582: Let $\omega_t(v)$ denote the indicator function for the
583: event that at time $t$ the site $v$ is white.
584:   Basically, all of the results 
585: concerning existence of exceptional times and lower bounds on Hausdorff 
586: dimension go through (with essentially the same proofs) in the more general setting 
587: where we assume that
588: 
589: \begin{enumerate}
590: \item[(i)] The processes $t\mapsto \omega_t(v)$ are independent (possibly with different
591: distributions depending on $v$) as $v$ runs over all sites. 
592: \item[(ii)] $\P[\omega_t(v)=1]=1/2$ for all $t$ and $v$. 
593: \item[(iii)] There is $c>0$ so that
594: $$
595: \bigl|\Eb{(-1)^{\omega_t(v)}(-1)^{\omega_s(v)}}\bigr| \le 1-c\, |t-s|
596: $$
597: for all $v$ and all $t$ and $s$ satisfying $|t-s|<c$.
598: \item[(iv)] For each $v$, the process $\omega_t(v)$  has
599: right continuous paths a.s. 
600: \end{enumerate}
601: (Condition (iv) is just a technical condition to insure that the events that we
602: consider are measurable.)
603: 
604: For results concerning upper bounds on Hausdorff dimension, the proofs go through assuming 
605: (i), (ii), (iv) and 
606: \begin{enumerate}
607: \item[(v)]  There is a $c>0$ so that
608: $$\E[\text{number of flips of }\omega_t(v) \text{ during }(t_1,t_2)] \le c (t_2-t_1)
609: $$
610: for all $v$ and all $t_1,t_2\in\R$ satisfying $t_1 < t_2<t_1+c$.
611: \end{enumerate}
612: However, for simplicity, we stick to the original setup.
613: 
614: \medskip
615: 
616: 
617: We mention a few other papers where analogous questions to those studied
618: in \cite{HPS} have been studied for other models. First, the results in
619: \cite{HPS} were extended and refined in \cite{PS}. Next, analogous questions
620: for the Boolean model where the points undergo independent Brownian motions was studied
621: in \cite{BMW}. Analogous questions for the lattice case for certain
622: interacting particle systems (where updates are not done in an independent fashion)
623: are studied in \cite{BrSt}. Finally in \cite{BeSc}, it is shown that
624: there are 
625: exceptional two dimensionl slices for the Boolean model in four dimensions.
626: 
627: The rest of the paper is organized as follows. 
628: In Section \ref{sec:dilute}, we will first provide background
629: on the Fourier-Walsh expansion of a function defined on $\{0,1\}^n$ as well
630: as connections with noise sensitivity and then continue on 
631: to give the proof of Theorem \ref{t.noise} as well as a 
632: generalization to the case where the algorithm is not required to always determine the value 
633: of the function $f$. (This will accomodate readers who are only interested in
634: Theorem \ref{t.noise}.) In Section \ref{sec:background}, we will give necessary background 
635: concerning percolation including a
636: discussion of critical site percolation on the triangular lattice as well as
637: a brief discussion of interfaces and critical exponents. 
638: In Section \ref{sec:percalg}, we will construct two 
639: algorithms determining certain events involving critical site percolation on the 
640: triangular lattice and analyze them to obtain upper bounds on the probability that 
641: a vertex is looked at during the algorithm. (For readers who only want to read
642: Theorem \ref{t.noise} and see how to apply it, they can just glance through 
643: Section \ref{sec:background} and then read Section \ref{sec:percalg}.)
644: Section~\ref{sec:percalg} also
645: gives a very detailed discussion of interfaces and
646: completes the proofs of Theorems \ref{th:crossingquantsquare} and 
647: \ref{th:crossingquant} by applying Theorem \ref{t.noise}. In 
648: Section \ref{sec:except}, we give the proof of Theorem \ref{th:except}, and
649: in Section \ref{sec:hd} we give the proof of Theorem \ref{th.hd}.
650: (Although the upper bound of $31/36$ given in Theorem \ref{th.hd} is a 
651: special case of Theorem \ref{th:cones}, we choose to give a
652: different direct proof of this 
653: without reference to the work done in Section \ref{sec:wedgesupper}.)
654: In Section \ref{sec:wedgeslower}, we prove the lower bounds 
655: on the Hausdorff dimension stated in
656: Theorems \ref{th:wedges} and \ref{th:cones}. 
657: In Section \ref{sec:wedgesupper}, we prove the upper 
658: bounds on the Hausdorff dimension
659: stated in Theorems \ref{th:wedges} and \ref{th:cones}. 
660: This will be based on a 
661: general formula which gives an upper bound on the 
662: Hausdorff dimension of various random sets (or proves they are empty)
663: in terms of {\sl influences} (Theorem~\ref{th.hdgeneral}).
664: We conclude the section by showing that
665:  Theorems \ref{th.no2clusters},
666: \ref{th.differentclusters}, \ref{th.halfplane} and \ref{th.halfplane2} 
667: immediately follow from Theorems \ref{th:wedges} and \ref{th:cones}.
668: After this, in Section \ref{sec:square},
669: we prove Theorem~\ref{th.no3clustersZ2}
670: and explain several plausible
671: ways in which the proof of Theorem \ref{th:except}
672: might be extended to bond percolation on the square lattice. 
673: In Section \ref{sec:questions}, we present some open questions.
674: 
675: Finally, the appendix proves some 
676: results about (non-dynamical) critical percolation
677: that are needed for Theorems~\ref{th.no2clusters}--\ref{th:cones}.
678: The main result is that if $r<r'<r''$, then the probability to have
679: $j$ crossings in a prescribed color sequence between distances $r$ and $r''$
680: from $0$ is equal to the product of the corresponding probabilities
681: between radii $r$ and $r'$ and between radii $r'$ and $r''$, times an error
682: term (depending on $j$) that is bounded away from $0$ and infinity.
683: Another consequence is that one gets good control on the positions
684: of the crossings at the inner and outer radii, as was already
685: demonstrated by Kesten~\cite[Lemma 2]{K2}. The proofs in the appendix
686: also establish the corresponding statements for critical bond percolation
687: in $\Z^2$.
688: 
689: 
690: 
691: \section{Noise sensitivity of algorithmically dilute functions}\label{sec:dilute}
692: 
693: 
694: In this section, we give some background and then prove Theorem \ref{t.noise}.
695: 
696: \subsection{Noise sensitivity background}\label{ss.noisebackground}
697: 
698: For a function $f$ from $\Omega=\Omega_n:=\{0,1\}^n$ to $\R$,
699: the Fourier-Walsh expansion of $f$ is given by
700: $ f = \sum_{S \subseteq [n]} \hat{f}(S) \chi_S,$ 
701: where, $\chi_S(T)=(-1)^{|S \cap T|}$ and $\hat{f}(S)= \int f \chi_S$.
702: Here and in the following, $\int$ refers to integration with respect to
703: uniform measure and we identify any vector 
704: $x\in\Omega_n$ with the subset $\{j\in[n]:x_j=1\}$ of
705: $[n]=\{1,2,\dots,n\}$.
706: Consequently, $|x|$ denotes the cardinality of that set; that is,
707: $|x|=\|x\|_1$ for $x\in\Omega_n$. The $\{\chi_S\}_{S\subseteq [n]}$ are an 
708: orthornormal basis for the $2^n$ dimensional vector space of functions from
709: $\Omega_n$ to $\R$. In particular,
710: $$
711: \|f\|^2=\sum_{S \subseteq [n]} \hat{f}(S)^2.
712: $$
713: 
714: We now generalize the definition of $N(A,\eps)$ given in the introduction
715: to any function $f:\Omega\to\R$ by defining
716: $$
717: N(f,\eps):=
718: \var\Bigl[\Eb{f(Y_1,\ldots,Y_{m})|X_1,\ldots,X_{m}}\Bigr].
719: $$
720: It is easy to see that (see page 14 in \cite{BKS}) 
721: \begin{equation}\label{e.page14}
722: N(f,\eps)= \sum_{\emptyset\neq S \subseteq [n]} \hat{f}(S)^2 (1-2\epsilon)^{2|S|}.
723: \end{equation}
724: This explains the importance of the Fourier-Walsh expansion in the study of
725: noise sensitivity.
726: 
727: \subsection{Proof of Theorem \ref{t.noise}}\label{ss.noisetheorem}
728: 
729: Before giving the proof, we discuss some heuristics.
730: One may first believe that an estimate such as~\eref{e.noise} would
731: be valid because when the algorithm terminates, the value of
732: $f$ is completely determined, and hence perhaps all the nonzero
733: Fourier coefficients $\hat f(S)\ne 0$, $S\ne\emptyset$, must
734: satisfy $S\cap \II\ne\emptyset$. However, this is easily shown not 
735: to be the case. At the $t$-th step of the algorithm,
736: after $t$ bits have been determined,
737: we may consider a new function $f_t$, which is $f$ 
738: with those determined bits substituted. 
739: If at the $(t+1)$-th step, the $i$-th bit $\omega_i$
740: of the input $\omega\in\Omega$ is examined, then
741: in the passage from $f_t$ to $f_{t+1}$, there is a collapsing
742: of Fourier coefficients: 
743: $\hat f_{t+1}(S)=\hat f_t(S)+(-1)^{\omega_i} \hat f_t(S\cup\{i\})$
744: and $\hat f_{t+1}(S\cup\{i\})=0$
745: for every $S\subseteq\nn\setminus\{i\}$.
746: Thus, the coefficient $\hat f_{t+1}(S)$ may vanish 
747: when some bit $i\in S$ is examined by time $t+1$ or when
748: some $i\notin S$ is chosen at time $t+1$ and it happens that
749: $\hat f_t(S)+(-1)^{\omega_i} \hat f_t(S\cup\{i\})=0$.
750: The latter, which we call \lq\lq collapsing from
751: above\rq\rq, may seem like a highly nongeneric 
752: situation. However, we cannot rule it out because
753: we are primarily interested in very non-generic functions,
754: namely, functions with values in $\{0,1\}$.
755: The proof below uses a simple decomposition argument to
756: handle the possibility of
757: collapsing from above.
758: 
759: In the following, we let $\tilde{\Omega}$ denote the probability
760: space that includes the randomness in the input bits of $f$ and the randomness
761: used to run the algorithm and we let $\E$ denote the
762: corresponding expectation.
763: Without loss of generality, elements of 
764: $\tilde{\Omega}$ can be represented as
765: $\tilde{\omega}=(\omega,\tau)$ where 
766: $\omega$ are the random bits and $\tau$ represents the
767: randomness necessary to run the algorithm.
768: 
769: \proof
770: Fix $k\ge 1$. Let 
771: $$
772: g(\omega):=\sum_{|S|=k} \hat f(S)\,\chi_S(\omega)\,,\qquad
773: \omega\in\Omega.
774: $$
775: The left hand side of~\eref{e.noise} is equal to $\|g\|^2$.
776: Let $\II\subset\nn$ be the random set of all bits examined by the algorithm.
777: Let $\ev A$ denote the 
778: minimal $\sigma$-field for which $\II$ is measurable and every
779: $\omega_i$, $i\in \II$, is measurable; this can be viewed as the
780: relevant information gathered by the algorithm.
781: For any function $h:\Omega\to\R$, let
782: $h_\II:\Omega\to\R$ denote the random function obtained by substituting
783: the values of the bits in $\II$. 
784: More precisely, if $\tilde{\omega}=(\omega,\tau)$
785: and $\omega'\in\Omega$, then $h_\II(\tilde{\omega})(\omega')$ is
786: $h(\omega'')$ where $\omega''$ is $\omega$ on $\II(\tilde{\omega})$ and is
787: $\omega'$ on $[n]\backslash \II(\tilde{\omega})$.
788: In this way, $h_\II$ is a random variable
789: on $\tilde{\Omega}$ taking values in the set of mappings from
790: $\Omega$ to $\R$ and it is immediate that this random variable is
791: $\ev A$-measurable.
792: When the algorithm terminates, the unexamined bits in $\Omega$
793: are unbiased and hence 
794: $\Eb{h\md \ev A}=\int h_\II(=\hat {h_\II}(\emptyset))$
795: where $\int$ is defined, as usual, to be integration with
796: respect to uniform measure on $\Omega$.
797: It follows that $\Es{h}=\Es{\int h_\II}$.
798: 
799: More generally, if $u:\R\to\R$, then
800: $(u\circ h)_\II=u\circ h_\II$ 
801: and hence, as above, $ \Es{u(h)}=\EB{\int u(h_\II)}$. 
802: In particular, for all $h$,
803: \begin{equation}\label{e.equiv}
804: \|h\|^2=\Eb{h^2}=\EB{\int h_\II^2}=\Eb{\|h_\II\|^2}.
805: \end{equation}
806: 
807: Since the algorithm determines  $f$, it is $\ev A$ measurable,
808: and we have
809: \begin{equation*}
810: \|g\|^2=\Es{g\,f}=\EB{\Eb{g\,f\md \ev A}}=
811: \EB{f\,\Eb{g\md\ev A}}.
812: \end{equation*}
813: Since $\Eb{g\md \ev A}=\hat {g_\II}(\emptyset)$,  Cauchy-Schwarz therefore gives
814: \begin{equation}\label{e.ept}
815: \|g\|^2 \le \sqrt{ \Es{\hat g_\II(\emptyset)^2}}\,\|f\|\,.
816: \end{equation}
817: We may write,
818: $$
819: \Es{\hat g_\II(\emptyset)^2}
820: =
821: \Eb{\|g_\II\|^2} -
822: \EB{\sum_{|S|>0} \hat g_\II(S)^2}.
823: $$
824: This and~\eref{e.equiv} with $h=g$ imply that
825: \begin{multline*}
826: \Es{\hat g_\II(\emptyset)^2}
827: \le
828: \|g\|^2
829: -\EB{\sum_{|S|=k}\hat g_\II(S)^2}
830:  \\
831: =
832: \sum_{S\subseteq\nn}\hat g(S)^2
833: -\EB{\sum_{|S|=k}\hat g_\II(S)^2}
834: =
835: \sum_{|S|=k}\Eb{\hat g(S)^2-\hat g_\II(S)^2}
836: \,.
837: \end{multline*}
838: It is easily seen that for any function $h$,
839: $h_\II=\sum_{S}\hat h(S)\,(\chi_S)_\II$.
840: We apply this with $h=g$.
841: Since $\hat g(S')=0$ if $|S'|> k$,
842: it follows that for all
843: $S\subset\nn$ satisfying $|S|=k$
844: $$
845: \hat g_\II(S)=
846: \begin{cases}
847: \hat g(S),& S\cap \II=\emptyset, \\
848: 0, &S\cap \II\ne\emptyset\,.
849: \end{cases}
850: $$
851: The above estimate for
852: $\Es{\hat g_\II(\emptyset)^2}$ therefore gives
853: $$%\begin{multline*}
854: \Es{\hat g_\II(\emptyset)^2}
855: \le
856: \sum_{|S|=k}\hat g(S)^2\,\Pb{S\cap \II\ne\emptyset}
857: \le \sum_{|S|=k}\hat g(S)^2\,\sum_{i\in S}\Ps{i\in \II}
858: \le \|g\|^2\,k\,\delta
859: \,.
860: $$%\end{multline*}
861: Substituting this estimate in~\eref{e.ept} and squaring the resulting
862: inequality completes the proof of the theorem.
863: \QED
864: 
865: 
866: The theorem may be easily generalized to situations where the
867: algorithm does not always determine the value of $f$ precisely;
868: that is, $f_\II(x)$ still depends on $x\in\Omega$.
869: 
870: Set 
871: $$
872: \var_\Omega(f_\II):=\int (f_\II)^2 - \Bigl(\int f_\II\Bigr)^2
873: =\sum_{S\ne\emptyset}\hat f_\II(S)^2\,,
874: $$
875: where the integrations are with respect to the uniform
876: probability measure on $\Omega$.
877: Note that $\Eb{\var_\Omega(f_\II)}$ is an indicator for how precisely
878: the algorithm can be used to approximate $f$;
879: when $\var_\Omega(f_\II)$ is small, with high conditional probability,
880: $|\hat f_\II(\emptyset)-f|$ is not too large.
881: 
882: When $\var_\Omega(f_\II)\ne0$, we have to replace the calculation
883: in the proof of Theorem~\ref{t.noise}
884: by the following
885: \begin{multline}
886: \|g\|^2=
887: \Eb{g_\II\,f_\II}
888: =\Eb{ g_\II\,\hat f_\II(\emptyset)}
889: + 
890: \Eb{g_\II\,(f_\II-\hat f_\II(\emptyset))}
891: \\
892: \le
893: \Bigl(\Es{\hat g_\II(\emptyset)^2}\,\Es{\hat f_\II(\emptyset)^2}\Bigr)^{1/2}
894: +
895: \Bigl(\Es{ g_\II^2}\,\Es{\var_\Omega( f_\II)}\Bigr)^{1/2}. 
896: \end{multline}
897: Since $\Eb{\hat f_\II(\emptyset)^2+\var_\Omega(f_\II)}=\Es{f^2}$,
898: we have $\Es{\hat f_\II(\emptyset)^2}\le \|f\|^2$.
899: Thus, the above gives
900: $$
901: \|g\|^2 \le
902: \sqrt{\Eb{\hat g_\II(\emptyset)^2}}\,\|f\|
903: +
904: \|g\|\,\sqrt{\Es{\var_\Omega(f_\II)}}\,.
905: $$
906: Using the same estimate for $\Eb{\hat g_\II(\emptyset)^2}$ as in the proof
907: of Theorem~\ref{t.noise}, we  obtain
908: $$
909: \|g\|^2
910: \le
911: \|g\|\,\|f\|\sqrt{k\,\delta}+
912: \|g\|\,\sqrt{\Es{\var_\Omega(f_\II)}}\,.
913: $$
914: Consequently, squaring both sides gives the following
915: generalization of~\eref{e.noise}:
916: \begin{multline}\label{e.gnoi}
917: \sum_{|S|=k}\hat f(S)^2
918: \le
919: \Bigl(\|f\|\sqrt{k\,\delta}+
920: \sqrt{\Es{\var_\Omega(f_\II)}}\Bigr)^2
921: \\
922: \le
923: 2\,k\,\delta\,\|f\|^2+
924: 2\,
925: \Es{\var_\Omega(f_\II)}\, .
926: \end{multline}
927: \bigskip
928: 
929: Theorem~\ref{t.noise} holds more generally than stated.
930: If $W$ is a random subset of $\nn$,
931: we say that $W$ is a {\bf witness} for $f:\Omega\to\R$ if
932: the value of $f$ is determined by its restriction to
933: $W$.  We say that $W$ is {\bf $\delta$-dilute} if
934: $\max_{i\in\nn}\Pb{i\in W(\omega)}\le\delta$.
935: The related notion of short witnesses is of central importance
936: in Talagrand's epic isoperimetric saga~\cite{Talagrand}.
937: The proof of Theorem~\ref{t.noise} holds in the more general setting where the
938: random
939: set $J$ is a witness
940: with the property that for all subsets $A\subseteq \nn$,
941: conditioned on $J=A$ (assuming this has positive probability)
942: and conditioned on $\omega$ restricted to $A$, the
943: $\{\omega_i\}_{i\not \in A}$ are uniform i.i.d.\ bits.
944: As pointed out to us by Asaf Nachmias, 
945: Theorem~\ref{t.noise} does not hold for arbitrary witnesses,
946: even if we allow for a multiplicative constant in the
947: right hand side of~\eref{e.noise}: if you take \lq\lq Recursive Ternary Majority\rq\rq\ on
948: $n=3^h$ bits, there is a (symmetric) witness having only $2^h$ elements,
949: yielding a $\delta$ which is $(2/3)^h$; however, the sum of the squares
950: of the level 1 Fourier coefficients is $(3/4)^h$.
951: 
952: \section{Percolation background and notations}\label{sec:background}
953: 
954: Duality plays a central role in the theory of percolation in two dimensions.
955: A dual-open path on the triangular grid is defined as a path
956: in the grid whose vertices are all closed. For the square grid,
957: a dual-open path is defined as a path in the dual of the square grid that
958: does not intersect any open edge in the primal grid.
959: The basic observation is that for site percolation on the 
960: triangular grid at $p=1/2$ the distribution of the collection of
961: dual-open paths is the same as the distribution of the collection of primal open paths.
962: (Sometimes, we use the term ``primal open path'', for an open path, to make
963: the distinction with the dual-open path clearer.)
964: For critical bond percolation on the square grid at $p=1/2$, the distribution
965: of the dual-open paths is the image of the distribution of the
966: open paths under translation by $(1/2,1/2)$.
967: This simple duality is one of the important ingredients in the proof by
968: Kesten that $p_c=1/2$ for these two percolation models~\cite[pg.~53]{Kesten}
969: and the earlier proof by Harris (see \cite{Harris})
970: that there is a.s.\ no infinite cluster at $p=1/2$.
971: 
972: For $0\le r <R<\infty$, let $A(r,R)$ denote the event that there is an open
973: crossing of the annulus $r\le|z|\le R$, namely, an open path connecting a vertex
974: inside the disk $|z|\le r$ to a vertex in $|z|\ge R$.
975: Let $\alpha(r,R)$ denote the probability of
976: ${A(r,R)}$ at percolation parameter $p=p_c=1/2$. 
977: Abbreviate $\alpha(0,R)$ by $\alpha(R)$.
978: For convenience, we adopt the convention $\alpha(r,R)=1$ whenever $r\ge R$.
979: The function $\alpha(r,R)$ is essentially multiplicative, in the following
980: sense: there is a constant $C>0$ such that for every $0\le r_1\le r_2\le r_3<\infty$,
981: \begin{equation}\label{e.multip}
982: C^{-1}\,\alpha(r_1,r_3)\le \alpha(r_1,r_2)\,\alpha(r_2,r_3)\le C\,\alpha(r_1,r_3)\,.
983: \end{equation}
984: In fact, this holds for critical bond percolation on the square grid as well
985: as for critical site percolation on the triangular grid.
986: The (standard) proof of~\eref{e.multip}
987: is based on the Harris-FKG inequality
988: %(see \cite{Grimmett}, page 34)
989: and the celebrated Russo-Seymour-Welsh (RSW) theorem
990: (see~\cite{Grimmett,Kesten}).
991: %(see \cite{Grimmett}, page 315).
992: A proof of a generalization of~\eref{e.multip} is given in the appendix.
993: Another consequence of RSW that we will use is the existence of a constant
994: $c>0$ such that for every $r>0$,
995: \begin{equation}\label{e.rsw}
996: c\le \alpha(r,2\,r)\,.
997: \end{equation}
998: 
999: The Stochastic L\"owner evolution (SLE) introduced in~\cite{Schramm} is a one parameter 
1000: family of random curves indexed by a real positive parameter $\kappa$.
1001: It was conjectured in \cite{Schramm} 
1002: that the scaling limit of outer boundaries of critical
1003: site percolation clusters on the triangular grid
1004: as well as
1005: bond percolation clusters on $\Z^2$ 
1006: are (chordal) SLE$_\kappa$ with ${\kappa=6}$.
1007: Smirnov~\cite{Smirnov,Smirnov1} proved the corresponding statement for critical site
1008: percolation on the triangular lattice. 
1009: (See also \cite{CN}.) We now explain some of this more precisely. We first
1010: perform independent site percolation on the upper half of the triangular lattice 
1011: but declare the sites $\{(k,0):k > 0\}$ to all be open and $\{(k,0):k\le 0\}$ to all be closed.
1012: In the hexagonal grid dual to the triangular grid, there will then be a
1013: unique path in the upper half plane from $(\frac{1}{2},0)$ to $\infty$ 
1014: which has white hexagons containing open 
1015: sites on the right and black hexagons containing closed
1016: sites on the left. (See Figure~\ref{f.pcurve}.)
1017: Smirnov's result is that the limit (in an appropriate toplogy) as the mesh size of the
1018: lattice goes to $0$ of the law of this path
1019: is chordal SLE$_6$.
1020: This path described above, which has open sites
1021: on its right and closed sites on its left, is an example of what is called an 
1022: {\it interface}.
1023: 
1024: \begin{figure}
1025: \centerline{\epsfysize=2.5in\epsfbox{pcurve.ps}}
1026: \begin{caption} {\label{f.pcurve}The percolation interface.}
1027: \end{caption}
1028: \end{figure}
1029: 
1030: The conformal invariance and the SLE description of critical percolation on 
1031: the triangular lattice allowed researchers to prove a number of conjectures by 
1032: physicists concerning so-called {\sl critical exponents}.
1033: For example, for critical site percolation on the triangular grid, it was established in 
1034: \cite{LSW} that 
1035: \begin{equation}\label{e.onearm}
1036: \alpha(R)=R^{-5/48 + o(1)} \text{ as $R\to\infty$}\,.
1037: \end{equation}
1038: In fact, the same proof actually gives for $R\ge r\ge 1$,
1039: \begin{equation}\label{e.onearmstrong}
1040: \alpha(r,R)=(R/r)^{-5/48 + o(1)} \text{ as $R/r\to\infty$}\,.
1041: \end{equation}
1042: % By a slight abuse of notations, here and in the following,
1043: % an expression of the form $s^{o(1)}$
1044: % will stand for any quantity $f$ such that there
1045: % is a function $g(s)$ of $s$ only satisfying
1046: % $g(s)^{-1}\le f\le g(s)$ and $\lim_{s\to L} \log g(s)/\log s=0$,
1047: % where $L=0+$ or $L=\infty$, depending on the context.
1048: 
1049: For $1\le r \le R$, the two arm function $\alpha_2(r,R)$ denotes the probability
1050: that there is both an open path from $|z|\le r$ to $|z|\ge R$ and also a 
1051: dual-open path from $|z|\le r$ to $|z|\ge R$. 
1052: We abbreviate $\alpha_2(1,R)$ by $\alpha_2(R)$.
1053: 
1054: Next, let $M$ be a half plane in $\R^2$ and let $v$ be some vertex in $M$
1055: satisfying $\dist (v,\p M)\le 2$.
1056: Denote by $\alpha^+(R)$ the probability that there is an open
1057: path in $M$ from $v$ to distance at least $R$ away from $v$.
1058: This quantity depends on $R$, $v$, and $M$, but the dependence on
1059: $v$ and $M$ will usually be suppressed.
1060: More generally, for $0\le r< R$, let 
1061: $\alpha^+(r,R)$ denote the probability that there is
1062: an open path in $M$ from some vertex $u$ satisfying $|u-v|\le r$ to some 
1063: vertex $w$ satisfying $|w-v|\ge R$.
1064: 
1065: As with $\alpha(r,R)$, we adopt the convention
1066: $\alpha^+(r,R)=1$ whenever $r\ge R$.
1067: 
1068: It is known that
1069: the functions $\alpha^+$ and $\alpha_2$ also satisfy~\eref{e.multip}, with possibly
1070: different constants. (When considering $\alpha^+$,
1071: this applies to any fixed choice of $M$ and $v$.)
1072: These inequalities are valid for site percolation on the triangular grid as well
1073: as bond percolation on the square grid. 
1074: A proof can be found in the appendix.
1075: For site percolation on the triangular grid, the corresponding exponents
1076: were established in~\cite{SW}:
1077: \begin{equation}\label{e.expo}
1078: \alpha^+(R)=R^{-1/3+o(1)},\qquad
1079: \alpha_2(R)=R^{-1/4+o(1)},
1080: \end{equation}
1081: as $R\to\infty$.
1082: In fact, the same proofs actually give for $R\ge r\ge 1$
1083: \begin{equation}\label{e.expostrong}
1084: \alpha^+(r,R)=(R/r)^{-1/3+o(1)},\qquad
1085: \alpha_2(r,R)=(R/r)^{-1/4+o(1)},
1086: \end{equation}
1087: as $R/r\to\infty$.
1088: 
1089: For bond percolation on the square grid, such exact estimates are unavailable,
1090: because there is currently no proof that the interface converges
1091: to SLE$_6$.
1092: In the case of the square grid, the
1093: estimate $\alpha(r,R)\le C\,(R/r)^{-\eps}$,
1094: where $C,\eps>0$ are constants, follows easily from the 
1095: RSW theorem (see \cite{Grimmett,Kesten}).
1096: The RSW proofs can give an actual value for $\eps$, but it is rather
1097: small. We can also use the obvious estimate $\alpha^+(r,R)\le\alpha(r,R)$
1098: to obtain a similar bound for $\alpha^+$.
1099: 
1100: \smallskip
1101: 
1102: We now give the precise definitions for wedges and cones.
1103: For this purpose, we first recall the definition of the infinitely
1104: branched cover of $\R^0$ over $0$.
1105: Let $X=\bl\{(z,\theta):z\in\C\setminus \{0\},\,\theta\in\R,e^{i\theta}|z|=z\br\}$,
1106: and set $\psi(z,\theta)=z$. On the surface $X$ we define the
1107: metric $d_X$ as the pullback of the Euclidean metric of $\R^2$ under
1108: $\psi$, namely, $d_X(x,y)$ is the infimum of the length
1109: of $\psi\circ \gamma$ for any continuous path $\gamma\subset X$ connecting $x$
1110: and $y$. Let $C_\infty$ denote the completion of $(X,d_X)$.
1111: Since $\R^2$ is complete, it is easy to see that
1112: $C_\infty\setminus X$ consists of a single point, which we denote by $0$.
1113: We extend the map $\psi$ by setting $\psi(0)=0$.
1114: Let $V$ be the set of points in $C_\infty$ that are mapped to vertices
1115: of the triangular grid under $\psi$. The triangular grid
1116: on $C_\infty$ has vertices $V$ and an edge between any two vertices
1117: at distance $1$ apart.
1118: Now the wedge $W_\theta\subset C_\infty$ is defined by
1119: $W_\theta:=\{0\}\cup \bl\{(z,\theta')\in X: \theta'\in[0,\theta)\br\}$.
1120: The triangular grid on $W_\theta$ is just the intersection of
1121: the triangular grid on $C_\infty$ with $W_\theta$.
1122: 
1123: On $C_\infty$ we may define the rotation $R_\theta$
1124: by $R_\theta(0)=0$ and $R_\theta(z,\theta')=(e^{i\theta}z,\theta+\theta')$.
1125: This is clearly an isometry of $C_\infty$.
1126: The cone $C_\theta$ is defined as the quotient $C_\infty/R_\theta$; that is,
1127: the set of equivalence classes of points in $C_\infty$, where two points
1128: are considered equivalent if one is mapped to the other by a power of
1129: $R_\theta$.  Now suppose that $\theta=n\,\pi/3$ where $n\in\N_+$. Then
1130: $R_\theta$ restricts to an isomorphism of the triangular grid on $C_\infty$.
1131: In this case we define the triangular grid on $C_\theta$ as the quotient
1132: of the grid on $C_\infty$ under $R_\theta$.
1133: In other words, the vertices are equivalence classes of vertices in
1134: $C_\infty$ and an edge appears between two equivalence classes if
1135: there is an edge connecting representatives of these classes.
1136: Note that $C_{2\pi}$ is just the euclidean plane with the triangular
1137: lattice.
1138: 
1139: \smallskip
1140: 
1141: 
1142: 
1143: We end this section with describing the so-called full and half plane 
1144: exponents for $k$-arm events that were derived in \cite{SW}.
1145: 
1146: For integer $k\ge 1$, let $A^k(r,R)$ be the event that there are $k$ disjoint
1147: crossings of the annulus $\{z\in\R^2: r\le |x|\le R\}$ with a specified color
1148: sequence (up to rotations), where we require that both colors appear 
1149: in the color sequence.
1150: For $k\ge 2$, and $r\ge 10k$, it was proved in \cite{SW} that
1151: \begin{equation}\label{e.fullpk}
1152: \alpha_k(r,R):=\Pb{A^k(r,R)}=(R/r)^{\frac{1-k^2}{12} + o(1)},
1153: \end{equation}
1154: as $R\to\infty$ while $r$ is fixed.
1155: (The result for $\alpha_2(R)$ in~\eref{e.expo} above is a special case of this.)
1156: Next, letting $A_+^k(r,R)$ be the event that there are $k$ disjoint paths in the 
1157: upper half plane from $|z|\le r$ to $|z|\ge R$ with any specified color sequence, then for
1158: $k\ge 1$, and $r\ge 10k$, it was proved in \cite{SW} that
1159: \begin{equation}\label{e.halfpk}
1160: \alpha^+_k(r,R):=\Pb{A_+^k(r,R)}=(R/r)^{\frac{-k(k+1)}{6} + o(1)},
1161: \end{equation}
1162: as $R\to\infty$ while $r$ is fixed. %
1163: (The result for $\alpha_+(R)$ in~\eref{e.expo} is a special case of this.) 
1164: Just as we said that the proofs of~\eref{e.expo} actually yield~\eref{e.expostrong},
1165: it is also the case that the proofs of~\ref{e.fullpk} and~\ref{e.halfpk} also
1166: yield versions when $R/r\to\infty$ while $r\ge 10\,k$ is not necessarily fixed.
1167: 
1168: 
1169: \section{Noise sensitivity for percolation} \label{sec:percalg}
1170: 
1171: \subsection{Simply connected case}\label{ss.simply}
1172: 
1173: To apply Theorem~\ref{t.noise} to percolation, we will need to describe algorithms achieving
1174: small revealment. One result of that nature is
1175: 
1176: 
1177: \begin{theorem}\label{t.square}
1178: Let $\cro=\cro_R$ be the indicator function for the 
1179: event that critical site percolation on the standard triangular grid
1180: contains a left to right crossing in some grid-approximating domain $D$
1181: to a large square of side length $R$.
1182: (For example, we could take $D$ to be the union
1183: of the hexagons in the dual grid that are contained in the
1184: square.)
1185: Then there is a randomized algorithm $A$ determining $\cro$ 
1186: such that $\delta_A \le R^{-1/4+o(1)}$
1187: as $R\to\infty$.
1188: 
1189: For critical bond percolation on the square grid, there is such an
1190: algorithm satisfying $\delta\le C\,R^{-a}$ for some constants
1191: $a,C>0$.
1192: \end{theorem}
1193: 
1194: \noindent
1195: Remark:
1196: Theorem~\ref{t.square} says that there is an algorithm for the
1197: relevant event which exposes on average at most $R^{7/4+o(1)}$ bits.
1198: Since the probability of points not too close to the boundary
1199: being pivotal is about $R^{-5/4+o(1)}$ (this is the 4 arm event)
1200: and for a monotone function $f$, $\hat f(\{i\})$ is
1201: the probability that $x_i$ is pivotal,
1202: the case $k=1$ in Theorem~\ref{t.noise} implies that the revealment
1203: is at least $R^{-1/2+o(1)}$.
1204: As pointed out in Peres, Schramm, Sheffield and Wilson~\cite{PSSW}, 
1205: this can also be obtained using an inequality of O'Donnell and Servedio.
1206: 
1207: Theorems~\ref{t.noise} and \ref{t.square} immediately give
1208: 
1209: \begin{corollary}\label{c.square}
1210: For every $\eps>0$ there is a constant $C=C(\eps)$ such that
1211: $$
1212: \sum_{|S|=k}\hat\cro_R(S)^2 \le C\, k\,R^{-1/4+\eps}
1213: $$
1214: holds for every $k=1,2,\dots$ and for every $R>0$.
1215: \QED
1216: \end{corollary}
1217: 
1218: The basic idea of the proof of Theorem~\ref{t.square} is rather simple.
1219: First we consider the interface started at the lower right corner and
1220: stopped when it hits the upper or left edges. (See Figure~\ref{f.interface}.)
1221: This interface is sufficient to determine $Q$.
1222: If we traverse the interface, revealing just the bits necessary for its
1223: determination, then with high probability most bits will not be
1224: examined. However, this does not yield an algorithm with small revealment
1225: because the hexagons near the lower right corner are very likely
1226: to be examined. 
1227: To rectify this problem, we instead start the interface at a different
1228: (random) location $p_0$ on the right side
1229: of $D$. 
1230: This determines the existence of a crossing from the right side above
1231: $p_0$ to the left side. Then another interface started at $p_0$ will
1232: determine if there is a crossing that starts below $p_0$.
1233: 
1234: \begin{figure}
1235: \centerline{\epsfysize=2in\epsfbox{sqpcurve.eps}}
1236: \begin{caption} {\label{f.interface}Following the interface from the corner.}
1237: \end{caption}
1238: \end{figure}
1239: 
1240: Let us now be a bit more precise regarding the notion of an interface.
1241: In the following, we use an equivalent dual version of the
1242: site percolation model on the triangular grid. The dual graph is
1243: the hexagonal grid, and we color the hexagon white if the site
1244: contained in it is open and black if the site in it is closed.
1245: Of course, there is no essential difference between these two
1246: representations. The advantage of this dual framework is that the
1247: figures are clearer and the notion of the interface is slightly more
1248: natural.
1249: 
1250: Note that bond percolation on the square grid also has a similar
1251: coloring representation. One such scheme is to
1252: color the squares of sidelength $1/2$ centered at the sites
1253: of $\Z^2$ white, color the squares of sidelength $1/2$ that are
1254: concentric with the square faces of $\Z^2$ black,
1255: and color each square of sidelength $1/2$ whose center is the
1256: midpoint of an edge of $\Z^2$ white or black if that edge is
1257: open or closed, respectively. See Figure~\ref{f.squareColorPerc}.
1258: This scheme has the important property that the boundary between
1259: the union of the white clusters is a $1$-manifold; that is,
1260: at every vertex of the grid $(1/2)\,\Z^2+(1/4,1/4)$ there are
1261: two edges that are on the common boundary.
1262: 
1263: \begin{figure}
1264: \centerline{\epsfysize=2.5in\epsfbox{sqPerc.eps}}
1265: \begin{caption} {\label{f.squareColorPerc}A color scheme for bond percolation on $\Z^2$.}
1266: \end{caption}
1267: \end{figure}
1268: 
1269: We now assume that $D$ is a bounded simply connected domain which is
1270: the interior of a union of hexagons in the hexagonal grid.
1271: Suppose that $p_0$ is a point on $\p D$ that is on the boundary
1272: of a single hexagon in $D$, and that $\zeta$ is a closed arc
1273: in $\p D\setminus\{p_0\}$.
1274: The interface in $\closure D$ from $p_0$ to $\zeta$ is a 
1275: random path $\beta$ contained in the $1$-skeleton of the hexagonal grid
1276: starting at $p_0$ and ending at a point in
1277: $\zeta$, as indicated in Figure~\ref{f.interfaceDef}.
1278: We can precisely define $\beta$ as the unique oriented simple path
1279: from $p_0$ to $\zeta$ that is contained in
1280: the union of the boundaries of the hexagons
1281: contained in $\closure D$ and satisfies
1282: (1) $\beta\cap\zeta$ consists of
1283: the terminal point of $\beta$,
1284: (2) whenever $\beta$ traverses an arc along the boundary
1285: of a black hexagon $H\subset\closure D$, the 
1286: arc is traversed counterclockwise around $\p H$, and
1287: (3) whenever $\beta$ traverses an arc along the boundary
1288: of a white hexagon $H\subset\closure D$, the 
1289: arc is traversed clockwise around $\p H$.
1290: It is easy to verify that this uniquely defines $\beta$,
1291: as follows.
1292: First, the initial arc of $\beta$ is determined by the
1293: color of the hexagon in $\closure D$ containing $p_0$.
1294: When $\beta$ first meets a hexagon contained in 
1295: $\closure D$, its turn is clearly specified.
1296: (If the hexagon is black, then $\beta$ must make
1297: a $\pi/3$ turn to the right,
1298: and if the hexagon is white, then $\beta$
1299: must make a $\pi/3$ turn to the left.)
1300: Now consider the situation in which $\beta$
1301: first meets a hexagon that is not contained in
1302: $\closure D$. Let $\beta'$ be the
1303: arc of $\beta$ from $p_0$ up to that point.
1304: Then $\beta$ at that point makes the turn into
1305: the component of $\closure D\setminus \beta'$
1306: that contains $\zeta$, as in the figure.
1307: 
1308: \begin{figure}
1309: \SetLabels
1310: \B(.56*.98)$\zeta$\\
1311: \R(.19*.48)$p_0$\\
1312: \endSetLabels
1313: %\ShowGrid
1314: \centerline{\epsfysize=2in%
1315: \AffixLabels{%
1316: \epsfbox{interfaceDef.eps}}%
1317: }
1318: \begin{caption} {\label{f.interfaceDef}An interface started at $p_0$
1319: and headed towards $\zeta$.}
1320: \end{caption}
1321: \end{figure}
1322: 
1323: Another way to describe this interface is that we color 
1324: the counterclockwise arc of $\p D$  from $p_0$ to
1325: $\zeta$ white and the clockwise arc from $p_0$ to
1326: $\zeta$ black, and $\beta$ then is the common boundary
1327: component between white and black in $\closure D$
1328: starting at $p_0$.
1329: We will call this type of interface a {\bf chordal interface},
1330: to differentiate it from the interface that we will later need
1331: when discussing the annulus crossing event.
1332: (The chordal interface was proved by Smirnov~\cite{Smirnov1} to converge to chordal
1333: SLE(6).)
1334: 
1335: 
1336: \proofof{Theorem~\ref{t.square}}
1337: We mostly concentrate on the case of site percolation on the
1338: triangular grid. The details in the case of bond percolation
1339: on $\Z^2$ are essentially the same.
1340: 
1341: The algorithm proceeds as follows. (See Figure~\ref{f.beta}.)
1342: There are four distinguished boundary arcs of $D$,
1343: which we call \lq\lq left\rq\rq, \lq\lq right\rq\rq, \lq\lq up\rq\rq\
1344: and \lq\lq down\rq\rq.
1345: Pick uniformly at random 
1346: an edge $e_0$ on the right hand boundary of 
1347: $D$, and let $p_0$ be its midpoint.
1348: Let $\zeta$ be the union of the top and left boundary
1349: segments of $D$.
1350: Explore the interface $\beta$ from $p_0$ to $\zeta$, examining the
1351: bits associated to sites in hexagons touching that interface,
1352: only as needed to continue with the determination of the interface.
1353: Note that the knowledge of $\beta$ suffices to determine if there
1354: is an open crossing from the right boundary of $D$ above $p_0$ to
1355: the left boundary of $D$: there is such a crossing if and only if
1356: $\beta$ terminates on the left boundary of $D$.
1357: 
1358: \begin{figure}
1359: \centerline{\epsfysize=2in%
1360: \SetLabels
1361: \L(1*.43)$\beta$\\
1362: \L(.95*.6)$p_0$\\
1363: \endSetLabels
1364: %\ShowGrid
1365: \AffixLabels{\epsfbox{beta.eps}}%
1366: \qquad
1367: \SetLabels
1368: \L(1.01*.25)$\beta'$\\
1369: \L(.95*.6)$p_0$\\
1370: \endSetLabels
1371: %\ShowGrid
1372: \epsfysize=2in%
1373: \AffixLabels{\epsfbox{betap.eps}}%
1374: }
1375: \begin{caption} {\label{f.beta}The interfaces $\beta$ and $\beta'$.}
1376: \end{caption}
1377: \end{figure}
1378: 
1379: Now let $\zeta'$ be the union of the bottom and left boundaries
1380: of $D$, and let $\beta'$ be the interface from $p_0$ to $\zeta'$
1381: that corresponds to the configuration $\omega'$ obtained by
1382: flipping all the colors of the hexagons in $\closure {D}$
1383: (alternatively, $\beta'$ is an interface that has black on the right
1384: and white on the left).
1385: Then $\beta'$ determines the existence of an open crossing from the
1386: right boundary below $p_0$ to the left boundary.
1387: Consequently, after the algorithm examines $\beta$ and $\beta'$,
1388: the correct value of $Q$ is determined.
1389: We now need to bound the revealment of this algorithm.
1390: 
1391: \begin{lemma}\label{l.Hdist}
1392: Let $\p_W$ denote the counterclockwise arc of $\p D\setminus\zeta$
1393: from $p_0$ to $\zeta$ (the white arc), and let
1394: $\p_B$ denote the clockwise arc of
1395: $\p D\setminus \zeta$ from $p_0$ to $\zeta$.
1396: Let $H$ be a grid hexagon in $\closure D$.
1397: Set $r_1:=\min\{\dist(H,\p_W),\dist(H,\p_B)\}$
1398: and $r_2:=\max\{\dist(H,\p_W),\dist(H,\p_B)\}$.
1399: Then
1400: $$
1401: \Pb{\p H\cap\beta\ne\emptyset \md p_0}
1402: \le \alpha_2(r_1)\,\alpha^+\bigl(2\,r_1+O(1),r_2-r_1-O(1)\bigr)\,.
1403: $$
1404: \end{lemma}
1405: \proof
1406: Suppose that $\p H\cap\beta\ne\emptyset$.
1407: Then there is a path in black hexagons from $\p H$
1408: to $\p_B$ and there is a path in white
1409: hexagons from $\p H$ to $\p_W$,
1410: because the chains of hexagons along the two
1411: sides of $\beta$ provide such paths.
1412: Suppose, for example, that
1413: $r_1=\dist(H,\p_W)$.
1414: Let $w$ be a closest point to $H$ in $\p_W$.
1415: Let $M$ be a half space that contains $D$,
1416: which satisfies $\dist(\p M,w)\le C$ where $C=O(1)$.
1417: (Here, we are using the fact that $D$ approximates a convex domain.)
1418: Let $w'$ be a point closest to $H$ on $\p M$.
1419: Then $\dist(w',H)\le r_1+C$.
1420: Set $R_1=2\,r_1+2\,C+2\diam(H)$ and $ R_2=r_2-r_1-C-\diam(H)$,
1421: and assume for now that $R_2>R_1$.
1422: Consider next the annulus centered at
1423: $w'$ with inner radius $R_1$ and outer radius $ R_2$.
1424: Now, $M$ intersected with this annulus
1425: contains a black crossing between the two boundary circles of this
1426: annulus (because there is a black crossing from
1427: $H$ to $\p_B$),
1428: and there are black and white crossings between
1429: $H$ and the circle of radius $r_1$ around the center
1430: of $H$.  These events are independent given $p_0$,
1431: which implies
1432: the lemma in the case $r_1=\dist(H,\p_W)$, $R_2>R_1$.
1433: If $R_2\le R_1$ and
1434: $r_1=\dist(H,\p_W)$, we only need to consider the
1435: crossings between $H$ and the circle of radius
1436: $r_1$ around its center.
1437: The case
1438: $r_1=\dist(H,\p_B)$ is treated similarly.
1439: \QED
1440: 
1441: \proofcont{Theorem~\ref{t.square}}
1442: Fix a hexagon $H\subset\closure D$.
1443: Note that the value of $r_1$ in the lemma does not depend on
1444: $p_0$, since $r_1=\dist(H,\p D\setminus\zeta)$.
1445: Let $z_0$ be the closest point to $H$ on the right boundary of
1446: the square which $D$ approximates.
1447: Observe that $r_2\ge |p_0-z_0|-O(1)$.
1448: This implies that for every $r\in[1,R]$, $\Pb{r/2\le r_2<r}\le O(r/R)$.
1449: Using the monotonicity of $\alpha^+$, Lemma~\ref{l.Hdist} therefore gives
1450: \begin{multline*}
1451: \Pb{\p H\cap\beta\ne\emptyset }
1452: \\
1453: \le O(1)\, \max_{0\le r_1\le R}\Bigl(
1454: \alpha_2(r_1)
1455: \sum_{j=0}^{\lceil \log_2 R\rceil}2^{-j}
1456: \alpha^+\bigl(2\,r_1+O(1),2^{-j}R-r_1-O(1)\bigr)
1457: \Bigr).
1458: \end{multline*}
1459: The same estimate also applies to $\beta'$, by symmetry.
1460: Consequently,~\eref{e.expostrong} gives
1461: \begin{multline*}
1462: \Pb{H\text{ examined by algorithm}}
1463: \\
1464: \le R^{o(1)}\max_{0\le r_1\le R}\Bigl(
1465: {r_1}^{-1/4}
1466: \sum_{j=0}^{\lceil \log_2 R\rceil}2^{-j}
1467: (2^{-j}R/r_1)^{-1/3}
1468: \Bigr)
1469: \\
1470: \le R^{o(1)}\max_{0\le r_1\le R}\bigl( {r_1}^{1/12} R^{-1/3} \bigr)
1471: = R^{-1/4+o(1)}\,,
1472: \end{multline*}
1473: as $R\to\infty$.
1474: This proves the theorem in the case of the triangular grid.
1475: 
1476: The proof for the square grid is essentially the same.
1477: Since in that case, we cannot use the values of the
1478: critical exponents, we just use the
1479: bounds $\alpha^+(r,r')\le C\,(r/r')^{\eps_0}$
1480: and $\alpha_2(r_1)\le C\, r_1^{-\eps_0}$, which are
1481: valid for some constants $C,\eps_0>0$.
1482: The theorem follows.
1483: \QED
1484: 
1485: 
1486: We now give the proof of Theorem~\ref{th:crossingquant}. We will not prove
1487: Theorem~\ref{th:crossingquantsquare}, but rather simply say that it is 
1488: proved in a similar way.
1489: 
1490: \proofof{Theorem~\ref{th:crossingquant}}
1491: Fix $\gamma < 1/8$.
1492: Let $f$ be the indicator function $f(\omega)=1_{A_m'}(\omega)$.
1493: By (\ref{e.page14}), we have
1494: $$
1495: N(A_m',m^{-\gamma})= \sum_{\emptyset\neq S \subseteq [n_m]} \hat{f}(S)^2 (1-2\,m^{-\gamma})^{2|S|}
1496: $$
1497: where $n_m$ is the number of sites in $D$.  By Corollary~\ref{c.square},
1498:  with $\epsilon >0$ chosen so that $2\,\gamma+\eps < 1/4$,
1499: this is at most
1500: \begin{multline*}
1501: C
1502: \sum_{k=1}^{n_m} (1-2m^{-\gamma})^{2k}\, k\,m^{-1/4+\eps}\\
1503: \le C\,{m^{-1/4+\epsilon}}
1504: \sum_{k=1}^{\infty}k\, (1-2\,m^{-\gamma})^{\frac{m^\gamma}{2}\frac{4k}{m^\gamma}} 
1505: \le C\,{m^{-1/4+\epsilon}}
1506: \sum_{k=1}^{\infty} k\, e^{-4k{m^{-\gamma}}} 
1507: \end{multline*}
1508: It is easy to check that
1509: $$
1510: \sum_{k=1}^{\infty} k\, e^{-4k{m^{-\gamma}}} \le O(m^{2\gamma})\,,
1511: $$
1512: and so the result follows immediately.
1513: \QED
1514: 
1515: \subsection{Annulus case}\label{ss.ann}
1516: 
1517: For the proof of Theorem~\ref{th:except}, we will need 
1518: the following variant of Theorem~\ref{t.square}
1519: regarding the percolation crossings of an annulus.
1520: 
1521: 
1522: \begin{theorem}\label{t.annulus}
1523: Let $2\le r< R$. 
1524: Let $\croRr$ be the indicator function for the 
1525: event that there is a crossing of the annulus
1526: $\{z\in\R^2: r\le z\le R\}$
1527: from the inner circle to the outer circle
1528: by a cluster of white hexagons.
1529: Then there is a randomized algorithm determining $\croRr$ with
1530: \begin{equation}\label{e.td}
1531: \delta\le r^{o(1)}\,\alpha(r,R)\,\alpha_2(r)\,.
1532: \end{equation}
1533: \end{theorem}
1534: 
1535: We stress that the $r^{o(1)}$ factor depends on $r$ only and not on $R$.
1536: It is possible to replace the factor $r^{o(1)}$ by $O(1)$,
1537: but in order to do this it seems that one would first need to
1538: appeal to the analogue of~\eref{e.multip} for $\alpha_2$, which
1539: is proved in the appendix.
1540: In order to have a more direct proof of our main theorem,
1541: we prefer, at this point, not to rely on the appendix.
1542: By~\eref{e.onearmstrong} and~\eref{e.expo},
1543: the right hand side in~\eref{e.td} is equal to
1544: $R^{-5/48+o(1)}\,r^{-7/48}$, but its writing in~\eref{e.td} is more
1545: suggestive and more useful.
1546: 
1547: 
1548: 
1549: \medskip
1550: 
1551: Since $\|\croRr\|^2=\alpha(r,R)$, 
1552: Theorems~\ref{t.noise} and~\ref{t.annulus} give 
1553: 
1554: \begin{corollary}\label{c.croRr}
1555: $$
1556: \sum_{|S|=k}\hat\croRr(S)^2
1557: \le 
1558: k\,r^{o(1)}\,\alpha(r,R)^2\,\alpha_2(r)
1559: $$
1560: holds when $1\le r<R<\infty$ and $k>0$.
1561: \QED
1562: \end{corollary}
1563: 
1564: Before we prove Theorem~\ref{t.annulus},
1565: we have to discuss the kind of interface that is used by the algorithm,
1566: as it is slightly different from the interface used to determine
1567: a possible crossing of a square.
1568: 
1569: Fix $R>0$ large.
1570: Let $\closure D=\closure D_R$
1571: be the union of the hexagons of the hexagonal grid 
1572: that intersect the disk $|z|\le R$.
1573: Let $V^*=V^*_R$ denote the set of vertices of the hexagonal grid
1574: that are in $\closure D$.
1575: Let $p_0$ be some point in $\p \closure D\setminus V^*$.
1576: Let $H_0$ denote the hexagon containing the origin, and
1577: let $q_0$ be some point in $\p H_0\setminus V^*$.
1578: We define the {\bf radial interface} $\beta=\beta(R,p_0,q_0,\omega)$,
1579: inductively as a simple path from $p_0$ to $q_0$.
1580: (See Figure~\ref{f.annulusinterface}.)
1581: The construction is segment by segment, and the concatenation
1582: of the first $m$ segments will be denoted $\beta_m$.
1583: If the (unique) hexagon in $\closure D$
1584: containing $p_0$ is white [respectively, black], then 
1585: the first segment $\beta_1$ of $\beta$ traverses the boundary
1586: of that hexagon [counter-] clockwise 
1587: until the first encounter with a point in $V^*$.
1588: Suppose inductively $\beta_m$ has been constructed,
1589: that it is a simple path, that
1590: $p_m\in V^*$ and $p_0$ are the two
1591: endpoints of $\beta_m$,
1592: and that there is a path $\alpha$ in
1593: the hexagonal grid in $\closure D$ from
1594: $p_m$ to $q_0$ whose only intersection with $\beta_m$ is $p_m$.
1595: The first step of such a path $\alpha$ is along an edge $\hat e$
1596: starting at $p_m$. If there is just one possible $\hat e$ among all
1597: such $\alpha$, then $\beta_{m+1}$ also uses that
1598: edge $\hat e$. Clearly, there are at most two possible $\hat e$,
1599: since the edge terminating at $p_m$ and used by $\beta_m$
1600: cannot be used. If there are two possible $\hat e$, then
1601: $\beta_{m+1}$ chooses between them according to the color of
1602: the hexagon containing them both; i.e., the hexagon just encountered
1603: by $\beta_m$. If that hexagon is white [respectively, black],
1604: then the edge chosen is the one that traverses $H$
1605: [counter-] clockwise. If the edge chosen contains $q_0$,
1606: then the path stops at $q_0$ and the construction terminates.
1607: Otherwise, $\beta_{m+1}$ is defined as the union of $\beta_m$ and
1608: the chosen continuation edge. This completes the definition of $\beta$.
1609: 
1610: \begin{figure}
1611: \SetLabels
1612: (.503*.50)$q_0$\\
1613: \T(.44*.01)$p_0$\\
1614: \B(.95*.61)$\beta^T$\\
1615: \B(.95*.43)$\beta\setminus\beta^T$\\
1616: \endSetLabels
1617: %\ShowGrid
1618: \centerline{\epsfysize=2.5in%
1619: \AffixLabels{%
1620: \epsfbox{annular.eps}}%
1621: }
1622: \begin{caption} {\label{f.annulusinterface}The radial interface
1623: $\beta$.}
1624: \end{caption}
1625: \end{figure}
1626: 
1627: 
1628: It is not hard to verify that for every simple path $\hat\beta$
1629: in the hexagonal grid from $p_0$ to $q_0$ that stays in
1630: $\closure D$, the probability that $\beta=\hat\beta$
1631: is precisely $2^{-n}$ if $n$ is the number of
1632: hexagons in $\closure D$ that intersect $\hat\beta$.
1633: However, we will not use this fact.
1634: 
1635: Let $r\in[0,R]$.
1636: We now define a truncated version of $\beta$, which will
1637: suffice, as we will see, to determine $\croRr$.
1638: We say that $\beta$ completed a [counter-] clockwise loop
1639: at some dual vertex $v\in V^*$ if $v$ is visited by $\beta$
1640: and there is a hexagon $H$ containing $v$ and  another  point
1641: $u\in\p H$, which was visited by $\beta$ prior to
1642: $v$, 
1643: and the oriented arc of $\beta$ from $u$ to $v$ together with the
1644: line segment $[v,u]\subset H$ form a [counter-] clockwise
1645: loop surrounding $0$.
1646: Let $\beta^T$ denote the initial segment of 
1647: $\beta$ up to the first time in which $\beta$ completed
1648: a counterclockwise loop around $0$ or until it hits
1649: $q_0$, if there is no such loop.
1650: 
1651: 
1652: \begin{lemma}\label{l.betadet}
1653: The truncated radial interface $\beta^T$ meets 
1654: the disk $|z|\le r$ if and only if $\croRr=1$.
1655: \end{lemma}
1656: \proof
1657: Let $[u,v]$ be an edge in $\beta^T$, with $u$ occuring
1658: before $v$ along $\beta$.
1659: We claim that if the hexagon $H$ immediately to the right
1660: of $[u,v]$ is contained in $\closure D$, then it is white.
1661: Indeed, suppose the contrary. Let $w$ be the first vertex along 
1662: $\beta$ in which $\p H$ is visited, and let
1663: $\beta^w$ be the initial segment of $\beta$ from $p_0$
1664: to $w$.  Observe that the counterclockwise arc from 
1665: $w$ to $v$ is a feasible continuation of $\beta^w$,
1666: since $\beta$ contains a path from $v$ to $q_0$ and
1667: there is no other point but $w$ in
1668: $\p H\cap\beta^w$.
1669: Since we are assuming that $H\subset\closure D$ is black,
1670: it follows that the immediate continuation of $\beta^w$
1671: was along $\p H$ in the counterclockwise direction until
1672: some $w'\in V^*$ is hit.
1673: 
1674: Consider the directed cycle obtained by joining the line segment
1675: $[u,w']$ to the arc of $\beta$ from $w'$ to $u$.
1676: This directed cycle surrounds $v$, because $w$ is connected
1677: in $\beta^w$ to $p_0$, which is certainly in the unbounded
1678: component of this cycle, and the line segment
1679: $[v,w]$ intersects the cycle precisely once,
1680: crossing the line segment $[u,w']$ inside $H$.
1681: Moreover, if we consider the orientation in which
1682: these two line segments cross, we conclude that
1683: the cycle surrounds $v$ counterclockwise.
1684: Because the arc of $\beta$ from $v$ to $q_0$
1685: does not intersect the cycle, we conclude that the cycle also
1686: surrounds $0$ counterclockwise.
1687: This contradicts the definition of the truncated path $\beta^T$,
1688: since we are assuming $v\in\beta^T$.
1689: This verifies our claim, that to the right of edges in
1690: $\beta^T$ are only white hexagons and hexagons that are not
1691: contained in $\closure D$.
1692: 
1693: Note that if $e$ and $e'$ are two consecutive segments
1694: along $\beta$, then the hexagon to the right of $e$ is
1695: either the same as the one to the right of $e'$, or these
1696: hexagons are adjacent. We therefore conclude that every
1697: white hexagon visited by $\beta^T$ is connected by
1698: a chain of white hexagons to $\p D$.
1699: Therefore, if $\beta^T$ hits the set $|z|\le r$,
1700: then clearly $\croRr=1$.
1701: 
1702: Now suppose that $\beta^T$ does not hit $|z|\le r$.
1703: This implies that $\beta^T$ has terminated by completing a
1704: counterclockwise loop around the set $|z|\le r$.
1705: Consider a hexagon $H$ on the inner boundary of this loop.
1706: Because the orientation of the loop is
1707: counterclockwise, the first time in which $H$ is
1708: visited, the path chose to traverse $\p H$ counterclockwise.
1709: This implies that the hexagon is black.
1710: Thus, there is a loop of black hexagons in $\closure D$
1711: that surrounds the set $|z|\le r$. This implies that $\croRr=0$.
1712: \QED
1713: 
1714: 
1715: We can now specify the algorithm promised by Theorem~\ref{t.annulus}.
1716: The algorithm starts by selecting the point $p_0$ uniformly along $\p\closure D$,
1717: and selecting $q_0$ arbitrarily in $\p H_0\setminus V^*$.
1718: It then proceeds to inspect the colors of the hexagons necessary to develop
1719: the truncated interface $\beta^T$, until it terminates or
1720: hits the set $|z|\le r$. At that point, the correct value
1721: of $\croRr$ is determined, by Lemma~\ref{l.betadet}.
1722: 
1723: In order to bound the revealment of this algorithm,
1724: it will be convenient to introduce a different interface, which in the end will
1725: turn out to be equivalent to $\beta$.
1726: 
1727: Let $\hat D$ denote the branched double cover of $\closure D$ about $0$,
1728: and let $\phi:\hat D\to \closure D$ denote the projection map.
1729: Concretely, define $\hat D$ as the preimage of
1730: $\closure D$ under the map $\phi(z)=z^2$.
1731: Let $\hat p_0$ be one of the preimages of $p_0$ under $\phi$,
1732: and let $\hat q_0$ be one of the preimages of $q_0$.
1733: Let $\hat H_0$ be the closure of one of the connected components
1734: of $\phi^{-1}(H_0)\setminus [\hat q_0,-\hat q_0]$.
1735: Let $\mfh$ denote the set of hexagons $H$ that are contained
1736: in $\closure D$.
1737: Let $\hmfh$ denote the set of connected components
1738: of preimages $\phi^{-1}(H)$, $H\in\mfh$, except that
1739: the single preimage of $H_0$ is replaced by
1740: the two sets $\hat H_0$ and $-\hat H_0$.
1741: Note that if $\hat H\in\hmfh$,
1742: then $-\hat H\in\hmfh$ and $\phi(\hat H)=\phi(-\hat H)\in\mfh$.
1743: Let $\hmfh' \subset \hmfh $ be a maximal collection
1744: of elements of $\hmfh$ with the property that
1745: $\hmfh'\cap \{-\hat H:\hat H\in\hmfh'\}=\emptyset$.
1746: Now color at random each of the elements of $\hmfh'$ white or
1747: black independently, with probability $1/2$.
1748: For every $\hat H\in\hmfh'$, let $-\hat H$ have the opposite
1749: color to the color of $\hat H$.
1750: 
1751: Now let $\hat \beta$ denote the chordal interface in $\hat D$
1752: from $\hat p_0$ to $-\hat p_0$, with white cells on the right
1753: and black cells on the left, as defined in the simply connected setting
1754: in Subsection~\ref{ss.simply}.
1755: That is, we consider the exterior of the counterclockwise arc from $\hat p_0$
1756: to $-\hat p_0$ along $\p \hat D$ as white, the exterior of the complementary
1757: arc as black, and take $\hat \beta$ as the interface between white and
1758: black passing through $\hat p_0$ and through $-\hat p_0$.
1759: Finally, let $\beta^\dagger:= \phi(\hat\beta)\setminus \interior(H_0)$.
1760: 
1761: \begin{lemma}
1762: Given $p_0$ and $q_0$,
1763: the laws of $\beta^\dagger$ and of $\beta$ are the same.
1764: \end{lemma}
1765: (In this statement, we consider $\beta$ as a set, and forget
1766: about the fact that it has the structure of an oriented path.)
1767: 
1768: \proof
1769: The map $z\mapsto -z$ preserves $\hat\beta$, by the symmetry
1770: of the interface. Consequently, near every point $p\in \beta^\dagger\setminus \{p_0,q_0\}$,
1771: $\beta^\dagger$ looks like a piecewise linear path. Moreover,
1772: $\beta^\dagger$ is connected and contains $p_0$. Since a compact simple path has two endpoints,
1773: we conclude that $q_0\in\beta^\dagger$ as well.
1774: We now consider building $\beta^\dagger$ by adding one segment at a time.
1775: When it hits a previously unvisited hexagon $H$ which is contained in $\closure D$,
1776: it is equally likely (given its past) to turn right or left.
1777: (This is because both preimages of $H$ are unvisited by both preimages of the
1778: past of $\beta^\dagger$.) When it hits a previously visited hexagon (or
1779: a hexagon that is not contained in $\closure D$), it turns in such a way
1780: that it will eventually be able to reach $q_0$ without crossing itself,
1781: and this uniquely specifies this turn. Consequently, the lemma follows.
1782: \QED
1783: 
1784: \begin{remark}
1785: The radial interface converges to radial SLE(6) as the mesh tends to zero.
1786: \end{remark}
1787: 
1788: 
1789: 
1790: 
1791: 
1792: \proofof{Theorem~\ref{t.annulus}}
1793: Given all of our preparations, the proof is rather easy.
1794: We have shown that the above algorithm provides the correct answer.
1795: It therefore remains to estimate its revealment.
1796: Consider some hexagon $H\subset\closure D$.
1797: We want to prove that the right hand side of~\eref{e.td}
1798: is an upper bound for the probability that $H$ is examined.
1799: Let $a:=\dist(0,H)$, $b:=\dist(H,\p\closure D)$ and $c:=\dist(p_0,H)$.
1800: Let $S_1$ be the disk of radius $(a\wedge b)/2$ concentric with $H$,
1801: and let $\hat S_1$ be one of the two connected components
1802: of $\phi^{-1}(S_1)$.
1803: We have to bound the probability that the algorithm inspects $H$.
1804: Clearly, we may assume $a\ge r-O(1)$.
1805: For $H$ to be inspected,
1806:  $\beta^T$ has to get to the circle
1807: $|z|=R\wedge (2\,a)$.  This probability
1808: is $\alpha(2\,a,R)$.
1809: Given that this has happened, how can we estimate the
1810: probability that $\beta$ is adjacent to $H$?
1811: At this point, we use the equivalence of
1812: $\beta$ and $\beta^\dagger$.
1813: The information that $\beta^\dagger$ reached the circle
1814: $|z|=R\wedge (2\,a)$ bears no impact on the distribution
1815: of the colors of the cells in $\hmfh$ whose images under
1816: $\phi$ intersect $S_1$. (Here, we assume that $a$ is not too small,
1817: so that the corresponding sets of cells are disjoint.
1818: Certainly $a>10$ suffices. If $a$ is smaller,
1819: then the estimate we are now striving for is trivial.)
1820: Since there is no hexagon intersecting both $\hat S_1$ and $-\hat S_1$,
1821: it follows that the conditional distribution of the
1822: colors of the cells meeting $\hat S_1$ is uniform
1823: i.i.d. Consequently, the conditional probability that
1824: $\beta^\dagger$ hits $H$ is bounded by $\alpha_2((a\wedge b)/2)$.
1825: Thus,
1826: $$
1827: \Pb{H\text{ visited}}\le
1828: O(1)\,\alpha(2\,a,R)\,
1829: \alpha_2((a\wedge b)/2)\,.
1830: $$
1831: In the case $b\ge a\ge 2\,r$, we may use independence on disjoint
1832: sets to conclude that 
1833: \begin{equation*}
1834: \begin{aligned}
1835: &\Pb{H\text{ visited}}
1836:  \\
1837: &\qquad
1838: \le
1839: O(1)\, \alpha_2(r)\,\alpha_2(2\,r,a/2)\,\alpha(2\,a,R)
1840: \le
1841: O(1)\,
1842: \alpha_2(r)\,\alpha(2\,r,a/2)\,\alpha(2\,a,R)
1843: \\
1844: &\qquad
1845: \overset{\eref{e.rsw}}{\le}
1846: O(1)\,
1847: \alpha_2(r)\,\alpha(r,2\,r)\,\alpha(2\,r,a/2)\,
1848: \alpha(a/2,2\,a)\,\alpha(2\,a,R)
1849: \\
1850: &\qquad
1851: \overset{\eref{e.multip}}{\le}
1852: O(1)\,
1853: \alpha_2(r)\,\alpha(r,R)\,.
1854: \end{aligned}
1855: \end{equation*}
1856: On the other hand, if $b\ge a$ and $a\le 2\,r$,
1857: then we use our assumption $a\ge r-O(1)$ and~\eref{e.expo}
1858: to get
1859: $\alpha_2(a/2)\le r^{-1/4+o(1)}\le\alpha_2(r)\,r^{o(1)}$,
1860: which is also sufficient.
1861: 
1862: In the case $b< a$, a similar argument 
1863: (and similar to the proof of Lemma~\ref{l.Hdist})
1864: shows that
1865: $$%\begin{multline*}
1866: \Pb{H\text{ visited }\md c}\le O(1)\,\alpha_2(b/2)\,\alpha^+\bl(2\,b+O(1),c-b-O(1)\br).
1867: $$%\end{multline*}
1868: Next, picking a constant $q\in(1/4,1/3)$, we then have by the above and~\eref{e.expostrong}
1869: $$%\begin{multline*}
1870: \Pb{H\text{ visited }\md c}
1871: \le O(1)\, \alpha_2(b/2)\, (c/b)^{-q}.
1872: $$%\end{multline*}
1873: As in the proof of Theorem~\ref{t.square}, we have $\Pb{2^j\le c<2^{j+1}} \le O(1)\,2^j/R$.
1874: It easily follows that
1875: $$
1876: \Pb{H\text{ visited}}\le
1877: O(1) \,\alpha_2(b/2)\, (R/b)^{-q}.
1878: $$
1879: If $b/2>r$, then we may estimate
1880: $$
1881: \alpha_2(b/2) \le \alpha_2(r)\, \alpha_2(2\,r,b/2) \le \alpha_2(r)\, \alpha(2\,r,b/2)
1882: $$
1883: and 
1884: $$
1885: (R/b)^{-q}
1886: \overset{\eref{e.expostrong}}{\le}
1887: O(1)\, \alpha(b/2,R)
1888: $$
1889: and we get from~\eref{e.multip} and the above
1890: $$
1891: \Pb{H\text{ visited}}\le O(1)\, \alpha_2(r)\, \alpha(r,R)\,.
1892: $$
1893: If $b/2\le r$, we use instead
1894: $$ 
1895: \alpha_2(b/2)\, (R/b)^{-q}\le O(1)\, \alpha_2(b/2)\, \alpha_2(b/2,r)\, \alpha(r,R)
1896: \overset{\eref{e.expostrong}}{\le}
1897: r^{o(1)}\, \alpha_2(r)\, \alpha(r,R)\,.
1898: $$
1899: This completes the proof.
1900: \QED
1901: 
1902: 
1903: \begin{remark}
1904: It is easy to see that if we assume the analogue of~\eref{e.multip} for 
1905: $\alpha_2$ proved in the appendix,
1906: then the $r^{o(1)}$ term in~\eref{e.td} can be replaced by $O(1)$.
1907: \end{remark}
1908: 
1909: \section{Exceptional times}\label{sec:except}
1910: 
1911: In this section we prove Theorem \ref{th:except}. 
1912: We point out that the absolute key necessary step is to get a good
1913: bound on the correlation for an event occurring at two different but
1914: close by times. Once this is done, the rest is fairly standard.
1915: Proposition A16 in Lawler~\cite{Lawler}
1916: indicates this general type of argument.
1917: 
1918: \proofof{Theorem~\ref{th:except}}
1919: By Kolmogorov's 0-1 law, it suffices to prove that with positive probability
1920: there are times in $[0,1]$ when the origin is in an infinite cluster.
1921: Fix $R>2$ large and let $V_{t,R}$ be the event
1922: that at time $t$ there is an open path from the origin to distance $R$ away.
1923: We then let
1924: $$
1925: X=X_R:= \int_0^1 1_{V_{t,R}} \, dt 
1926: $$
1927: be the Lebesgue measure of the set of times in $[0,1]$ at which $V_{t,R}$ occurs.
1928: The first moment of $X$ is given by
1929: $$
1930: \Eb{X}=\int_0^1 \Pb{V_{t,R}}\,dt=\Pb{V_{0,R}} =\alpha(R)\,.
1931: $$
1932: The second moment is
1933: \begin{equation}\label{e.2nd}%\begin{multline*}
1934: \Eb{X^2}=
1935: \EB{\int_0^1\int_0^1 1_{V_{s,R}}\,1_{V_{s',R}}\,ds\,ds'}
1936: =
1937: \int_0^1\int_0^1 \Pb{V_{s,R}\cap V_{s',R}}\,ds\,ds'.
1938: \end{equation}%\end{multline*}
1939: 
1940: 
1941: For each site $v$ we let 
1942: $$
1943: \chi_v^s:=\begin{cases} -1&v \text{ is open at time } s\\
1944: 1&\text{otherwise},
1945: \end{cases}
1946: $$
1947: and for a finite set of sites $S$ set
1948: $$
1949: \chi_S^s:=\prod_{v\in S}\chi_v^s\,.
1950: $$
1951: Fix $s,s'\in[0,1]$, and set $t:=|s-s'|$.
1952: Recall that the state of a site $v$ flips between closed and open
1953: with rate $1/2$. Equivalently, we may think of the state as being
1954: re-randomized with rate $1$.
1955: Consequently, $\Pb{\chi_v^{s'}=\chi_v^s \md \omega_s} = e^{-t}+(1-e^{-t})/2
1956: =(1+e^{-t})/2$,
1957: and hence,
1958: $$
1959: \Eb{\chi_v^s\,\chi_v^{s'}}=\exp(-t)
1960: \,,\qquad
1961: \Eb{\chi_S^{s}\,\chi_S^{s'}}=\prod_{v\in S}\Eb{\chi_v^s\,\chi_v^{s'}}=\exp(-t\,|S|)\,.
1962: $$
1963: Moreover, if $S\ne S'$, then $\Eb{\chi_S^s\,\chi_{S'}^{s'}}=0$.
1964: Consequently, if $f$ is a function depending on the states of finitely many lattice
1965: points and has the expansion
1966: $f(\omega)=\sum_S\hat f(S)\,\chi_S(\omega)$, then
1967: \begin{equation}\label{e.timeexpansion}
1968: \Eb{f(\omega_s)\,f(\omega_{s'})}=
1969: \sum_S\hat f(S)^2 \exp(-t\,|S|)\,.
1970: \end{equation}
1971: 
1972: Let $f_r^R(\omega)$ be as in Theorem~\ref{t.annulus}.
1973: Fix some $t\in (0,1]$ and let $r\in [2,R)$.
1974: Clearly, $0\le f_0^R(\omega)\le f_0^r(\omega)\,f_{2r}^R(\omega)$ for every $\omega$.
1975: Consequently,
1976: \begin{multline*}
1977: \Pb{\Vss}=
1978: \Eb{f_0^R(\omega_s)\,f_0^R(\omega_{s'})}
1979: \le
1980: \Eb{f_0^r(\omega_s)\,f_{2r}^R(\omega_s)\,
1981: f_0^r(\omega_{s'})\,f_{2r}^R(\omega_{s'})}
1982: \\
1983: =
1984: \Eb{f_0^r(\omega_s) \, f_0^r(\omega_{s'})}
1985: \Eb{ f_{2r}^R(\omega_s)\,f_{2r}^R(\omega_{s'})}
1986: \le
1987: \Eb{f_0^r(\omega_s) } \Eb{ f_{2r}^R(\omega_s)\,f_{2r}^R(\omega_{s'})} .
1988: \end{multline*}
1989: (To obtain the second equality, we have used the independence on 
1990: disjoint sets of sites.) Thus,
1991: $$
1992: \Pb{\Vss}\le
1993: \alpha(r)\,
1994: \Eb{ f_{2r}^R(\omega_s)\,f_{2r}^R(\omega_{s'})}
1995: =\alpha(r)\, \sum_S e^{-t|S|}\hat f_{2r}^R(S)^2.
1996: $$
1997: The latter sum restricted to $S$ with $|S|=k$ for fixed $k\ne 0$
1998: is estimated using Corollary~\ref{c.croRr},
1999: while for $k=0$, we use $\hat f_{2r}^R(\emptyset)=\alpha(2r,R)$. This yields
2000: $$
2001: \Pb{\Vss}\le
2002: \alpha(r)\,\Bigl(\alpha(2\,r,R)^2+
2003: r^{o(1)}\,\sum_{k=1}^\infty 
2004: e^{-kt}
2005: k\,\alpha(2\,r,R)^2\, \alpha_2(r)\Bigr).
2006: $$
2007: It is easy to check that $\sum_{k=1}^\infty k\,e^{-kt}\le O(t^{-2})$.
2008: This and the inequalities~\eref{e.multip} and~\eref{e.rsw} allow
2009: us to write this estimate as
2010: \begin{equation}\label{e.zz}
2011: \Pb{\Vss}\le
2012: O(1)\,
2013: \alpha(R)^2\,\alpha(r)^{-1}\bl(1+r^{o(1)}\,t^{-2}\,\alpha_2(r)\br)\,.
2014: \end{equation}
2015: We proved the above claim for all 
2016: $t\in (0,1]$ and $r\in [0,R)$ but
2017: now we observe that~\eref{e.zz} is also trivially true
2018: when $r \ge R$ as well. 
2019: We now choose $r=2\,t^{-8}=2\,|s-s'|^{-8}$.
2020: Applying this in~\eref{e.zz} with~\eref{e.onearm} and~\eref{e.expo} gives
2021: \begin{equation}\label{e.uselater}
2022: \Pb{V_{s,R}\cap V_{s',R}}
2023: \le O(1)\,\alpha(R)^2\, |s-s'|^{-5/6+o(1)}.
2024: \end{equation}
2025: Hence,
2026: \begin{equation}\label{e.fin}
2027: \int_0^1\int_0^1\Pb{V_{s,R}\cap V_{s',R}}\,ds\,ds'
2028: \le O(1)\,\alpha(R)^2.
2029: \end{equation}
2030: 
2031: The Cauchy-Schwarz inequality tells us that
2032: $$
2033: \Pb{X>0} \ge 
2034: \frac{
2035: \Eb{X}^2
2036: }{
2037: \Eb{X^2}
2038: }\,.
2039: $$
2040: Consequently, the above inequality, the fact that $\Eb{X}=\alpha(R)$,
2041: the expression~\eref{e.2nd} for $\Eb{X^2}$ and~\eref{e.fin}
2042: show that $\inf_{R>0}\Pb{X_R>0}>0$.
2043: Let $T_R:=\{t\in[0,1]:V_{t,R}\text{ holds}\}$.
2044:  Fatou's lemma tells us that with positive probability
2045: $T_R\ne \emptyset$ for infinitely many $R\in\N$.
2046: Since $T_R\supset T_{R'}$ when $R'>R$,
2047: this implies that 
2048: $$
2049: \Pb{\cap_{R>0}\,\, \{T_R\ne\emptyset\}}>0.
2050: $$
2051: Our goal is to show that $\Pb{\bigcap_R T_R \neq \emptyset}>0$.
2052: Since the $T_R$'s are not closed sets,
2053: $\cap_{R>0}\,\, \{T_R\ne\emptyset\}$
2054: does not immediately imply $\bigcap_R T_R\ne\emptyset$.
2055: (The reason that $T_R$ is not necessarily closed is that the set of times at
2056: which an edge is open is not a closed set since we have a right continuous
2057: process.) This technicality is taken care of by the following lemma from \cite{HPS}. 
2058: 
2059: \begin{lemma}\label{l.HPS} (\cite{HPS})
2060: Let $0<p<1$ and let $G$ be any graph where $\pi_p(\calC)=0$.
2061: Let $\{\omega_t\}$ represent our dynamical percolation process in
2062: that $\omega_t(v)$ is the state of vertex $v$ at time $t$.
2063: Consider the process $\{\bar\omega_t\}$ obtained from $\{\omega_t\}$
2064: by setting, for every vertex $v$,
2065: the set  $\{t\, : \, \bar\omega_t(v) =1\}$ to be 
2066: the closure of the set $\{t\, : \, \omega_t(v) =1\}$.
2067: Then $\bPsi_p$-a.s., for every vertex $v$ we have
2068: $$
2069:   \{t \in [0, \infty) \: :\:  v \, 
2070: \mbox{ \rm percolates in } \bar\omega_t \} \, = \,
2071:  \{t \in [0, \infty) \: :\:  v \, \mbox{ \rm percolates in } \omega_t \} \,.
2072: $$
2073: In particular, a.s.\ this set of times is closed.
2074: \end{lemma}
2075: 
2076: Returning to our proof,
2077: let $\overline{T_R}$ be the closure of $T_R$. It is easily checked that
2078: $$
2079: \bigcap_{R>0}\overline{T_R}= 
2080:   \{t \in [0,1] \: :\:  0 \, 
2081: \mbox{ \rm percolates in } \bar\omega_t \},
2082: $$
2083: where $\{\bar\omega_t \}$ is defined in Lemma \ref{l.HPS}.
2084: By compactness, if the $T_R$'s are all nonempty, it follows that 
2085: $\bigcap_R\overline{T_R}$ is nonempty. This implies that there 
2086: is some time at which $\bar\omega_t$ percolates and hence by
2087: Lemma \ref{l.HPS}, some time at which the original process $\omega_t$ percolates.
2088: \QED
2089: 
2090: For future reference, we note that Lemma~\ref{l.HPS} implies that a.s.
2091: \begin{equation}\label{e.ctr}
2092: {\bigcap_{R>0}\overline{T_R}= 
2093: \bigcap_{R>0}{T_R} }.
2094: \end{equation}
2095: 
2096: 
2097: \section{Hausdorff dimension of exceptional times} \label{sec:hd}
2098: 
2099: In this section, we prove Theorem \ref{th.hd}. This is separated into two theorems,
2100: Theorem \ref{t.lowerbound} and Theorem \ref{t.upperbound}, where lower and upper bounds
2101: are given. We point out however that the lower bound is simply a refinement of
2102: the argument for proving that there exist exceptional times.
2103: First note that the fact that the  Hausdorff dimension is an almost sure
2104: constant follows immediately from ergodicity.
2105: 
2106: \begin{theorem}\label{t.lowerbound}
2107: A.s.,  the Hausdorff dimension of the set of exceptional times is at least $1/6$.
2108: \end{theorem}
2109: 
2110: \proof
2111: Fix $\gamma < 1/6$. It suffices by ergodicity and countable additivity
2112: to show that with positive probability,
2113: the set of exceptional times in $[0,1]$ at which the origin percolates
2114: has Hausdorff dimension at least $\gamma$.
2115: For each integer $R$, let as before $V_{t,R}$ be the event
2116: that at time $t$ there is a path from the origin to distance $R$ away and
2117: define a random measure $\sigma_R$ on $[0,1]$ by
2118: $$
2119: \sigma_R(S)=\frac{1}{\alpha(R)}\int_{S}1_{V_{t,R}}dt
2120: $$
2121: for each Borel set $S\subset [0,1]$.
2122: 
2123: The results in the previous section immediately give that
2124: $\E[\| \sigma_R \|] =1$ and $\E[\| \sigma_R \|^2] \le O(1)$
2125: where $\| \sigma_R \|$ denotes the total variation of the measure $\sigma_R$.
2126: 
2127: Cauchy-Schwarz gives
2128: $$
2129: \Eb{\|\sigma_R\|^2}^{1/2}\,\Pb{\|\sigma_R\|>1/2}^{1/2}
2130: \ge
2131: \Eb{\|\sigma_R\| 1_{\|\sigma_R\|>1/2}}
2132: \ge
2133: \Es{\|\sigma_R\| }-1/2=1/2.
2134: $$
2135: Consequently, 
2136: $\Pb{\| \sigma_R \| > 1/2} \ge  C_1 $
2137: for some constant $C_1 >0$.
2138: Given a measure $m$ on $[0,1]$ and $\gamma >0$, let
2139: $$
2140: {\cal E}_{\gamma} (m) = \int \!\!\int {|t-s|}^{- \gamma} \, dm (t) \,   dm(s) .
2141: $$
2142: Note that
2143: $$
2144: \Eb{{\cal E}_{\gamma} (\sigma_R)}
2145: =
2146: \EB
2147: {\int_0^1\!\int_0^1\frac{d\sigma_R (t) \,   d\sigma_R(s)}{ {|t-s|}^ \gamma}  }=
2148: \int_0^1\!\int_0^1
2149: \frac{\Pb{{V_{t,R}}\cap{V_{s,R}} }}{{\alpha(R)^2}\, |t-s|^\gamma}\, dt\,ds \,.
2150: $$
2151: Therefore, 
2152: by~\eref{e.uselater} and $\gamma<1/6$,
2153: $$
2154: C_2:=\sup_R
2155: \Eb{{\cal E}_{\gamma} (\sigma_R)}<\infty\,.
2156: $$
2157: 
2158: By Markov's inequality, for all $R$ and for all $T$,
2159: $$
2160: \Pb{{\cal E}_{\gamma} (\sigma_R)\ge  C_2T} \le 1/T.
2161: $$
2162: Choose $T$ so that $1/T < C_1/2$.
2163: Letting 
2164: $$
2165: U_R=\{\| \sigma_R \| > 1/2\}\cap \{{\cal E}_{\gamma} (\sigma_R)\le  C_2T\},
2166: $$
2167: by the choice of $T$, we have that 
2168: $$
2169: \Pb{U_R} \ge C_1/2.
2170: $$
2171: By Fatou's lemma,
2172: $$
2173: \Pb{\limsup_{R\to\infty} U_R} \ge C_1/2.
2174: $$
2175: We now show that on the event $\limsup_R U_R$, the Hausdorff dimension of the 
2176: set of percolating times in $[0,1]$ is at least $\gamma$. 
2177: Let $\overline{T_R}$ again be the closure of the set of times in $[0,1]$ at which 
2178: there is a path from the origin to distance $R$ away. Clearly $\sigma_R$ is supported on
2179: $\overline{T_R}$.
2180: By~\eref{e.ctr}, it suffices to prove that
2181: $\bigcap_{R>0}\overline{T_R}$ has Hausdorff dimension at least $\gamma$
2182: on the event $\limsup_R U_R$.
2183: This is achieved in the following
2184: (deterministic) lemma, which completes the proof.
2185: \QED
2186: 
2187: \begin {lemma}\label{l.deterministic}
2188: Let $D_1\supseteq D_2 \supseteq D_3 \ldots$
2189: be a decreasing
2190: sequence of compact subsets of $[0,1]$,
2191: and let $\mu_1,\mu_2,\dots$ be a sequence of positive measures
2192: with $\mu_n$ supported on $D_n$.
2193: Suppose that there is a constant $C$ such that for infinitely many
2194: values of $n$, we have
2195: \begin{equation}\label{e.2cond}
2196: \| \mu_n \| > 1/C,\quad \text{ and } \quad{\cal E}_{\gamma} (\mu_n)\le C.
2197: \end{equation}
2198: Then the Hausdorff dimension of $\bigcap_n D_n$ is at least $\gamma$. 
2199: \end {lemma}
2200: 
2201: \proof
2202: Choose a sequence of integers $\{n_k\}$ for which
2203: (\ref{e.2cond}) holds.
2204: Note that $\|\mu_{n_k}\|^2\le \mathcal E_\gamma(\mu_{n_k})\le C$.
2205:  By compactness, 
2206: choose a further subsequence $\{n'_{k}\}$ of $\{n_k\}$ so that
2207: $\mu_{n'_{k}}$ converges weakly to some positive measure $\mu_{\infty}$.
2208: Clearly $\mu_{\infty}$ is supported on $\bigcap_n D_n$ and
2209: $\| \mu_\infty \| \ge 1/C$. For all $M$, we have that
2210: \begin{multline*}
2211: \int \!\!\int {|x-y|}^{- \gamma} \wedge M \, d\mu_\infty (x) \,   d\mu_\infty(y) =
2212: \\ \lim_{k\to\infty}
2213: \int \!\!\int {|x-y|}^{- \gamma} \wedge M \, d\mu_{n'_k} (x) \,   d\mu_{n'_k}(y) \le C.
2214: \end{multline*}
2215: Now let $M\to\infty$ and apply the monotone convergence theorem to conclude that
2216: $$
2217: \int \!\!\int {|x-y|}^{- \gamma}\, d\mu_\infty (x) \,   d\mu_\infty(y) 
2218: \le C.
2219: $$
2220: Since $\| \mu_\infty \| > 0$, it now follows from Frostman's theorem
2221: (see for example,~\cite{Kahane})
2222: %(see for example, \cite{Kahane}, page 133)
2223: that the Hausdorff dimension of $\bigcap_n D_n$ is at least $\gamma$. 
2224: \QED
2225: 
2226: \begin{theorem}\label{t.upperbound}
2227: A.s., 
2228: the Hausdorff dimension of the set of exceptional times is at most $31/36$.
2229: \end{theorem}
2230: 
2231: \proof
2232: Let $U_n$ be the event that there is a time in $[0,1/n]$ for which
2233: the origin percolates. Since the set of vertices which are open for some $t\in [0,1/n]$
2234: is an i.i.d.\ process with density $1/2 +(1-e^{-1/(2n)})/2\le 1/2+1/n$, it is immedate that
2235: $$
2236: \Pb{U_n}\le \pi_{\frac{1}{2}+\frac{1}{n}} (\calC_0) ,
2237: $$
2238: where $\calC_0$ is the event that the origin percolates.
2239: By page 3 of \cite{SW}, for every $\epsilon >0$, there is a $C$ so that
2240: \begin{equation}\label{e.offcrit}
2241: \pi_{\frac{1}{2}+\frac{1}{n}} (\calC_0)\le 
2242: C\, n^{\eps-5/36}.
2243: \end{equation}
2244: Now let 
2245: $$
2246: N_n=\sum_{j=1}^n 1_{U_{j,n}}\,,
2247: $$
2248: where $U_{j,n}$ is the event that there is a time in $[(j-1)/n,j/n]$ for which
2249: the origin percolates (so $U_{1,n}=U_n$ above). 
2250: By the above, we have that $\Eb{N_n}\le C n^{\frac{31}{36}+\epsilon}$.
2251: It follows that
2252: $$
2253: \lim_n \frac{\Eb{N_n}}{n^{\frac{31}{36}+2\epsilon}}=0
2254: $$
2255: and so from Fatou's lemma, we get
2256: $$
2257: \EB{\liminf_n \frac{N_n}{n^{\frac{31}{36}+2\epsilon}}}=0.
2258: $$
2259: Therefore
2260: $$
2261: \liminf_n \frac{N_n}{n^{\frac{31}{36}+2\epsilon}}=0
2262: $$
2263: a.s. This says that a.s.\ for infinitely many $n$, the set of exceptional
2264: times in $[0,1]$ at which the origin percolates
2265: can be covered by $n^{\frac{31}{36}+2\epsilon}$ intervals of length $1/n$.
2266: Hence, the Hausdorff dimension of the set of these exceptional times is 
2267: at most $\frac{31}{36}+2\epsilon$ a.s. By countable additivity, we are done.
2268: \QED
2269: 
2270: \begin{remark}
2271: The upper bound will be proved again by a different argument when
2272: we prove Theorem~\eref{th:cones}.
2273: The above proof is included 
2274: here, because it is shorter.
2275:  One should nonetheless
2276: point out that the above argument uses~\eref{e.offcrit},
2277: while the argument below is more self contained.
2278: \end{remark}
2279: 
2280: \section{Exceptional times for $k$-arm events} \label{sec:wedgeslower}
2281: 
2282: 
2283: In this section, we give the proofs of the lower bounds in
2284: Theorems \ref{th:wedges} and \ref{th:cones},
2285: but generally omit those details which are the same as in the
2286: corresponding proofs of Theorems~\ref{th:except} and~\ref{t.lowerbound}.
2287: 
2288: For $\theta>0$ and integer $k\ge 1$, let
2289: $A_{W_\theta}^k(r,R)$ be the event that
2290: we have $k$ disjoint crossings of alternating colors (with black most clockwise)
2291: between distances $r$ and $R$ of the origin in $W_\theta$
2292: and let $\alpha_{W_\theta}^k(r,R)=\Pb{A_{W_\theta}^k(r,R)}$. 
2293: If $r$ is suppressed, then it is taken to be $10\,k$.
2294: 
2295: We will, of course, need the asymptotics of $\alpha_{W_\theta}^k(r,R)$.
2296: For this purpose, conformal invariance will be used.
2297: Although when $\theta>2\pi$ the surface
2298: $W_\theta$ is not planar and conformal invariance is usually stated for planar domains,
2299: the proof of conformal invariance certainly holds in this setting.
2300: The asymptotic decay as $R/r\to\infty$ of the probability of $k$ disjoint crossings
2301: between distances $r$ and $R$ in $W_{\theta}$ from the origin in 
2302: the percolation scaling limit is determined using conformal invariance.
2303: Specifically, the map $z\mapsto z^{\pi/\theta}$ maps $W_{\theta}$ to the
2304: upper half plane, and we may conclude from~\eref{e.halfpk}
2305: that the decay (for the percolation
2306: scaling limit)  is of the form $(R/r)^{\frac{-\pi k(k+1)}{6\theta}+o(1)}$,
2307: as $R/r\to\infty$ while $k$ stays fixed.
2308: Then, one can conclude, as for the other exponents we have discussed,
2309: that for $R\ge r\ge 10\,k$,
2310: \begin{equation}\label{e.awt}
2311: \alpha_{W_\theta}^k(r,R)= (R/r)^{\frac{-\pi k(k+1)}{6\,\theta}+o(1)},
2312: \end{equation}
2313: as $R/r\to\infty$ while $k$ is fixed, using the argument given in~\cite{SW}. 
2314: We will also use the fact that the quasi-multiplicativity relations~\eref{e.multip}
2315: and~\eref{e.rsw} hold for $\alpha_{W_\theta}^k$ and
2316: for $\alpha_2$. This is proved in the appendix;
2317: see Remark~\ref{r.qmc}.
2318: 
2319: 
2320: \proofof{lower bound in Theorem~\ref{th:wedges}}
2321: We first handle the case $k=1$ and therefore abbreviate temporarily
2322: $\alpha_{W_\theta}^1(r,R)$ by $\alpha_{W_\theta}(r,R)$.
2323: (A different approach will be needed for $k\ge 2$.)
2324: Let $D$ be the union of the hexagons in $W_\theta$ that contain
2325: points whose distance from the origin is in $[r,R]$.
2326: Let $\p_R D$ and $\p _r D$ denote the set of points in
2327: $\p D$ that are at distance $\ge R$ [respectively,
2328: $\le r$] from the origin.
2329: Also, we denote by $\p^0 D$ and $\p^\theta D$,
2330: the components of $\p D\cap\p W_\theta$ that are at angle
2331: about $0$ [respectively, about $\theta$] in radial coordinates on $W_\theta$.
2332: 
2333: The algorithm we use to determine if
2334: there exists a crossing of $D$ is essentially the same as
2335: the algorithm determining the existence of a left to right crossing of a square,
2336: where $\p_R D$
2337: plays the role of the right side of the square and
2338: where $\p_r D$
2339: plays the role of the left side of the square.
2340: (This is of course crucial; if we reversed things, then the hexagons
2341: near the inner circle would be revealed with too high of a probability.)
2342: It is clear that this algorithm works and so we now need
2343: to compute its revealment.
2344: We will show that the revealment is
2345: \begin{equation}\label{e.revealment}
2346: O(1)\,\alpha_2(r)\,\alpha_{W_\theta}(r,R)\,.
2347: \end{equation}
2348: Using~\eref{e.awt} and~\eref{e.expostrong}, one can show that this is
2349: essentially (i.e., up to some $O(1)$ factor) monotone decreasing
2350: in $r$ in the relevant range $\theta>8\,\pi/3$.
2351: 
2352: We just look at the first interface arising in the algorithm,
2353: the one which terminates when it hits $\p_r D\cup\p^\theta D$,
2354: since the estimates for the second interface will be essentially the same.
2355: 
2356: Fix some hexagon $H\subset D$.
2357: Let $s=\dist(H,\p_R D\cup \p^0 D\cup \{0\})$ with $0$ denoting the origin.
2358: We also use $|p|$ to denote distance from $0$ in $W_\theta$.
2359: We distinguish several different cases.
2360: 
2361: \medskip\noindent
2362: {\bf Case 1}: $\dist(H,\{0\})=s$. 
2363: For $H$ to be visited, we need our 2 arms event holding within distance $s/2$ of $H$
2364: and a crossing of the desired color between distance $2s$ and distance $R$ from the 
2365: origin. These are independent and we get that $H$ is visited with probability 
2366: at most $\alpha_2(s/2)\,\alpha_{W_\theta}(2s,R)$.
2367: By the analogues of~\eref{e.multip} and~\eref{e.rsw} for
2368: $\alpha_2$ and $\alpha_{W_\theta}$, this is compatible with
2369: our claimed revealment~\eref{e.revealment}.
2370: 
2371: 
2372: \medskip\noindent
2373: {\bf Case 2}: $\dist(H,\p_R D)=s$. 
2374: As in the proof of Theorem~\ref{t.annulus} with $c:=\dist(p_0,H)\wedge (R/2)$,
2375: we obtain
2376: $$
2377: \Pb{H\text{ visited }\md p_0}\le O(1)\,\alpha^+(2s+O(1),c-s-O(1))\,\alpha_2(s/2).
2378: $$
2379: Proceeding as in that proof, this is also compatible with
2380: our claimed revealment~\eref{e.revealment}.
2381: 
2382: \medskip\noindent
2383: {\bf Case 3}: $\dist(H,\p^0 D)=s$. 
2384: Let $w\in \p^0 D$ so that $\dist(H,w)=s$. We separate Case 3 into 3 subcases.
2385: 
2386: \medskip\noindent
2387: {Case 3(a)}: $s \ge |w|/2$.
2388: Then the triangle inequality gives $\dist(H,0)\le 3s$. 
2389: For $H$ to be visited, we need our 2 arms event holding within distance $s/2$ of 
2390: $H$ and a path of the desired type between distance $4s$ and distance $R$ 
2391: from the origin. These are independent and we get that $H$ is visited with 
2392: probability at most $\alpha_2(s/2)\,\alpha_{W_\theta}(3s,R)$,
2393: which is compatible with our claimed revealment~\ref{e.revealment}.
2394: 
2395: \medskip\noindent
2396: {Case 3(b)}: $s \le |w|/2 \le R/4$.
2397: For $H$ to be visited, we need our 2 arms event holding within distance $s/2$ of $H$,
2398: a white crossing in the half annulus centered at $w$ with outer radius $|w|$ and inner radius $2s$
2399: (which is identical to a half-annulus in a half-plane;
2400: $|w|\le R/2$ guarantees that the above half-annulus does not intersect $\p_R D$)
2401: and a white crossing between distance $2|w|$ and distance $R$ from the 
2402: origin.
2403: These are independent and we
2404: get that $H$ is visited with probability at most 
2405: $$
2406: \alpha_2(s/2)\,\alpha_{+}(2s,|w|)\, \alpha_{W_\theta}(2|w|,R).
2407: $$
2408: Since up to an $O(1)$ factor $\alpha_2(s)\,\alpha_{+}(s,|w|)$ is increasing in $s$,
2409: the product of the first two terms is at most
2410: $O(1)\,\alpha_2(|w|)$ and since $|w|\ge 2\,s\ge r$, the whole product is at most 
2411: $$
2412: O(1)\,\alpha_2(r)\,\alpha_{W_\theta}(r,R).
2413: $$
2414: 
2415: \medskip\noindent
2416: {Case 3(c)}: $|w| \ge R/2$; $s\le |w|/2$.
2417: For $H$ to be visited, we need our 2 arms event holding 
2418: within distance $s/2$ of $H$ and if 
2419: $2s<d(p_0,w)$ it is also necessary that
2420: a white crossing 
2421: occurs between distances $2s$ and $d(p_0,w)\wedge |w|$ from
2422: $w$.
2423: (Note that the latter event takes place in the upper half plane.)
2424: These are independent and since $|w|\ge R/2$ we get 
2425: $$
2426: \Pb{H\text{ visited }\md p_0}\le O(1)\,\alpha^+\bl(2s,d(p_0,w)\wedge (R/2)\br)\,\alpha_2(s/2)\,.
2427: $$
2428: As in Case 2, this is compatible with~\eref{e.revealment}.
2429: 
2430: \medskip
2431: This covers all possible cases, and hence establishes that the
2432: revealment is as claimed.
2433: 
2434: \medskip
2435: 
2436: We now proceed to discuss the algorithm and the revealment when $k>1$.
2437: It turns out simplest in fact to
2438: modify the event $A_{W_\theta}^k(r,R)$ as follows. Partition  the
2439: outer boundary $\p_R D$ into $k$ arcs  of roughly equal diameter
2440: $Y_1,Y_2,\ldots,Y_k$ (ordered counterclockwise) and let $\tilde{A}_{W_\theta}^k(r,R)$
2441: be the event that for every odd [respectively, even] $i\in\{1,2,\dots,k\}$ there is a black
2442: [respectively, white]
2443: crossing in $D$ from $\p_r D$ to $Y_i$.
2444: Thus, instead of looking at the set of times for which $A_{W_\theta}^k(r_0,R)$
2445: occurs (where $r_0=10\,k$, say), we will look at the set of times at which $\tilde A_{W_\theta}^k(r_0,R)$
2446: occurs.
2447: Clearly, $\tilde{A}_{W_\theta}^k(r,R)\subset A_{W_\theta}^k(r,R)$,
2448: and therefore this is justified.
2449: We will also use the relation
2450: \begin{equation}\label{e.AA}
2451: \Pb{{A}_{W_\theta}^k(r,R)} \le C^\theta_k\, \Pb{ \tilde A_{W_\theta}^k(r,R)}\,,
2452: \end{equation}
2453: for some constant $C_k^\theta$, depending only on $k$ and $\theta$,
2454: which holds by Remark~\ref{r.primus}.
2455: 
2456: If $Y\subset \p_R D$ is an arc, let $A_Y^1(r,R)$
2457: [respectively, $A_Y^{-1}(r,R)$] be the
2458: event that there is a white [repectively, black] crossing from $Y$ to $\p_r D$
2459: in $D$.
2460: Suppose that each for each $i=1,2,\dots,k$, we have
2461:  a partition $Y_i=Y_{i,+}\cup Y_{i,-}$
2462: of $Y_i$ into two arcs $Y_{i,+}$ and $Y_{i,-}$.
2463: Since $A_{Y_i}^{\pm 1}(r,R)=A_{Y_{i,+}}^{\pm 1}(r,R)\cup A_{Y_{i,-}}^{\pm 1}(r,R)$,
2464: we have
2465: \begin{equation}\label{e.union}
2466: \tilde{A}_{W_\theta}^k(r,R)
2467: =\bigcup_{y\in \{-,+\}^k}
2468: \bigcap_{i=1}^k A_{Y_{i,y_i}}^{{(-1)}^i}(r,R)\,.
2469: \end{equation}
2470: 
2471: The algorithm starts out by picking points $x_i\in Y_i$,
2472: randomly, uniformly and independently.
2473: Then $Y_{i,+}$ and $Y_{i,-}$ are chosen as the two components
2474: of $Y_i\setminus \{x_i\}$.
2475: For each of the $2^k$ possible $y\in\{-,+\}^k$,
2476: the algorithm then proceeds to determine if
2477: the corresponding component
2478: $$
2479: A(y):=
2480: \bigcap_{i=1}^k A_{Y_{i,y_i}}^{{(-1)}^i}(r,R)
2481: $$
2482: of~\eref{e.union} has occured.
2483: For that purpose, interfaces are started at each of the points $x_i$,
2484: and are followed until the event has been determined one way or the other.
2485: (Of course, the interface will have either white on the left and
2486: black on the right or vice versa, depending on the color of crossing
2487: it is meant to detect and whether the corresponding arc
2488: $Y_{i,\pm}$ is to the left or right of $x_i$.)
2489: However, the order in which the interfaces are extended is somewhat important.
2490: A simple rule that works is that among the hexagons that are necessary
2491: to extend the $k$ interfaces one more step, the algorithm chooses the one that is
2492: farthest away from $0$.
2493: The event $A(y)$ is decided positively only if
2494: all $k$ interfaces reach $\p_r D$.
2495: 
2496: The revealment of this algorithm is at most $k\,2^k$ times the maximum
2497: probability that the interface started at $x_i$ visits a hexagon
2498: $H$ before the determination of the corresponding
2499: $A(y)$ is terminated. Here, the maximum is over all hexagons $H\subset D$
2500: and all $i\in\{1,2,\dots,k\}$.
2501: The corresponding bound is attained as in the case
2502: $k=1$, but now $\alpha_{W_\theta}^1$ is replaced by
2503: $\alpha_{W_\theta}^k$. 
2504: Our rule of thumb for selecting which interface to
2505: extend guarantees that we never examine a hexagon
2506: $H$ unless $\tilde A_{W_\theta}^k\bl(\dist(0,H)+O(1),R\br)$ occurred.
2507: As in the case $k=1$, when estimating the revealment it is important that
2508: $\alpha_2(r,R)\le O(1)\,\alpha_{W_\theta}^k(r,R)$.
2509: In the range $\theta> 4\,\pi\,k\,(k+1)/3$, which is the
2510: relevant range for the lower bound in Theorem~\ref{th:wedges}, this follows
2511: from~\eref{e.awt}.
2512: 
2513: The remainder of the proof goes through as before.
2514: \QED
2515: 
2516: \proofof{lower bound in Theorem~\ref{th:cones}}
2517: Here we simply say that the proof for the lower bounds in 
2518: Theorem~\ref{th:wedges} can be carried out in a similar way. 
2519: In fact, for $k\ge 2$, the proof
2520: is simpler topologically than the $k=1$ case for the plane,
2521: since we do not need to worry about interfaces making complete circuits
2522: around the origin 
2523: (if this ever happens, the event in question cannot occur and we stop the 
2524: algorithm).
2525: \QED
2526: 
2527: \section{Upper bounds for $k$-arm times} \label{sec:wedgesupper}
2528: 
2529: The following result,
2530: which will be useful for the proofs
2531: of the upper bounds in Theorems~\ref{th:wedges} and~\ref{th:cones},
2532: is abstract: the graph structure does not play any role.
2533: Let $A$ be an event involving independent Bernoulli $(1/2,1/2)$ random variables
2534: $X_1,X_2,\ldots,X_m$. Recall that the influence of the index $i$ on $A$, denoted $I_i(A)$,
2535: is the probability that $X_i$ is pivotal; namely,
2536: that changing the value of $X_i$ changes whether $A$ 
2537: occurs or not. 
2538: The sum of the influences is denoted by $I(A)=\sum_i I_i(A)$.
2539: 
2540: 
2541: 
2542: \begin{theorem} \label{th.hdgeneral}
2543: Let $\{A_n\}_{n\ge 1}$ be some sequence of events 
2544: in $\{0,1\}^V$, each depending on only finitely many 
2545: coordinates. 
2546: Assume that $\lim_{n\to\infty} \Ps{A_n} = 0$.
2547: Let $\omega_t$ be the Markov process on $\{0,1\}^V$ where
2548: independently $0$'s go to $1$ at rate $1/2$,
2549: $1$'s go to $0$ at rate $1/2$ and started according
2550: to its stationary distribution $\pi_{\frac{1}{2}}$.
2551: Let $T$ be the set of exceptional times $t$
2552: at which $\omega_t\in\bigcap_{n\ge 1}A_n$.
2553: If $\liminf_{n\to\infty} I(A_n) < \infty$, then 
2554: $T=\emptyset$ a.s. Otherwise, the Hausdorff dimension of $T$ is a.s.\ at most
2555: \begin{equation}\label{e.expression}
2556: \liminf_{n\to\infty}
2557: \left({{1-\frac{\log \Ps{A_n}}{\log I(A_n)}}}\right)^{-1}\,.
2558: \end{equation}
2559: \end{theorem} 
2560: 
2561: 
2562: \proof
2563: Let $T_n:=\{t\in [0,1]:\omega_t\in A_n\}$, let $\partial T_n$ be the boundary
2564: points of $T_n$ in $(0,1)$ and set $N_n:=|\partial T_n|$. We claim that
2565: \begin{equation}\label{e.Nn=In} 
2566: \Eb{N_n}={I(A_n)}/{2}.
2567: \end{equation}
2568: To see this, write
2569: $N_n=\sum_v N^v_n$ where $N^v_n$ counts the number of elements in
2570: $\partial T_n$ at which time the vertex $v$ flipped. We now need to
2571: show that, for each vertex $v$, $\Eb{N^v_n}$ is $I_v(A_n)/2$.
2572: Given a time interval $[t,t+dt]$, the probability that 
2573: there is a time point in the interval which contributes to $N^v_n$
2574: is equal to $I_v(A_n)\,dt/2 + o(dt)$ and the probability of $k\ge 2$ such time
2575: points is clearly $O(dt^k)$. From this,~(\ref{e.Nn=In}) easily follows.
2576: 
2577: For any $\eps > 0$, let $T^\eps_n$ be the $\eps$-neighborhood of
2578: $T_n$ intersected with $[0,1]$.
2579: Since $T^\eps_n\subseteq T_n\cup \bigcup_{x\in \p T_n}[x-\eps,x+\eps]$,
2580: \begin{equation}\label{e.star} 
2581: \mu(T^\eps_n)\le \mu(T_n)+ 2\,N_n\,\eps\,,
2582: \end{equation}
2583: where $\mu$ denotes Lebesgue measure.
2584: For any set $U$ and $\epsilon>0$, let $\calN(U,\epsilon)$ denote the number of 
2585: $\epsilon$ intervals needed to cover $U$.
2586: From the above, using the fact that the intervals comprising
2587: $T^\eps_n$ all have length at least $\eps$, it follows that
2588: $\calN(T^\eps_n,\eps)\le2\,\mu(T^\eps_n)\,\eps^{-1}$, and so,
2589: using~\ref{e.star},
2590: $$
2591: \calN(T_n,\eps)\le \calN(T^\eps_n,\eps)\le
2592: {2\,\mu(T_n)\,}{\eps^{-1}}+ 4\,N_n\,.
2593: $$
2594: Therefore, by Fubini's theorem and~\ref{e.Nn=In}, 
2595: \begin{equation}\label{e.covernumber}
2596: \Eb{\calN(T_n,\eps)}\le {2\,\Ps{A_n}}\,{\eps^{-1}}+ 2\,I(A_n)\,.
2597: \end{equation}
2598: 
2599: We now temporarily assume that $\liminf_{n\to\infty} I(A_n) =\infty$. 
2600: Let $a_n=\Pb{A_n}/I(A_n)$, which goes to $0$ as $n\to\infty$.
2601: By (\ref{e.covernumber}), we have 
2602: \begin{equation}\label{e.neededbound}
2603:   \Eb{\calN(T,a_n)}\le 
2604:   \Eb{\calN(T_n,a_n)}\le 4\,I(A_n)\,.
2605: \end{equation}
2606: By passing to a subsequence if necessary, we assume with no loss of generality
2607: that the $\liminf$ in~\eref{e.expression} is a limit. Let $L$ denote the value
2608: of that limit.
2609: It is elementary to check that for every $\eps>0$, for all sufficiently large $n$,
2610: $$
2611: I(A_{n})\le \left(\frac{I(A_{n})}{\Ps{A_{n}}}\right)^{L+\epsilon}.
2612: $$
2613: This together with (\ref{e.neededbound})
2614: implies that  %
2615: the Hausdorff dimension of $T$ is at most $L+\epsilon$ a.s. As 
2616: $\epsilon$ is arbitrary, this completes the proof in the case
2617: $I(A_n)\to\infty$.
2618: 
2619: Since $T_n\ne \emptyset$ implies that $N_n\ge 1$ or $T_n\supseteq (0,1)$,
2620: it follows by (\ref{e.Nn=In}) and Markov's inequality that 
2621: $$
2622: \Pb{T_{n} \neq \emptyset}\le \Pb{A_{n}} + I(A_{n})\,.
2623: $$
2624: Thus, $T=\emptyset$ a.s.\ when $\liminf_n I(A_n)= 0$.
2625: 
2626: The case $\liminf_n I(A_n)\in (0,\infty)$ requires a different argument.
2627: Let $\eps_n=\sqrt{\Ps{A_n}}$. By (\ref{e.covernumber}), we have 
2628: $
2629: \liminf_{n\to \infty}\Eb{\calN(T_n,\eps_n)} < \infty
2630: $.
2631: Since $\eps_n\to 0$, the cardinality $|T|$ of $T$ is bounded from above by
2632: $ \liminf_{n\to\infty}\calN(T_n,\eps_n) $.
2633: Fatou's lemma yields $\Eb{|T|}< \infty$ and
2634: hence $\Pb{|T|<\infty}=1$. We finally conclude that
2635: $\Ps{T\neq\emptyset}=0$ by combining
2636: \cite[Theorem 6.7]{GS} and \cite[(2.9)]{FG}.
2637: \QED
2638: 
2639: \proofof{Theorem~\ref{th:cones}}
2640: Since the lower bounds have been established in Section~\ref{sec:wedgeslower},
2641: it remains to prove the upper bounds.
2642: Fix $k=1$ or $k>1$ even.
2643: Let $A_R$ be the event that there are $k$ disjoint
2644: crossings of the annulus $D_R:=\{z\in C_\theta: 10k\le |z|\le R\}$, where we require
2645: that the colors be alternating if $k\ne 1$.
2646: Here, $|z|$ denotes the distance to $0$, which is the apex of the cone $C_\theta$.
2647: One can prove that 
2648: \begin{equation}\label{e.coneexpo}
2649: \Pb{A_R}
2650: =\begin{cases}
2651: R^{-5\pi/(24\theta)+o(1)}&k=1,\\
2652: R^{{(1-k^2)\pi}/({6\theta}) + o(1)}& k>1,
2653: \end{cases}
2654: \end{equation}
2655: in the very same way that we have justified~\eref{e.awt}.
2656: By Theorem~\ref{th.hdgeneral} (and easy algebraic manipulation), it therefore suffices to prove
2657: that
2658: \begin{equation}\label{e.pivotal}
2659: I(A_R)/\Ps{A_R}\le R^{3/4+o(1)}.
2660: \end{equation}
2661: Let $H$ be a hexagon in $C_\theta$, and let $s=s(H)$ be the distance from
2662: $H$ to $0$.
2663: For $H$ to be pivotal 
2664: it is necessary that there would be
2665: $k$ disjoint (alternating, if $k>1$)
2666: crossings from distance $10\,k$ to
2667: $s/2$ from the origin (unless $s/2\le 10\,k$)
2668: and between distances $2\,s$ and
2669: $R$ (unless $2\,s\ge R$).
2670: Likewise, there should be $4$ alternating crossings
2671: between $H$ and distance
2672: $(s/2)\wedge\dist(H,\p D_R)$ from $H$.
2673: These events are independent.
2674: Using the quasi-multiplicative property
2675: of the $k$-arm crossing events (Remark~\ref{r.qmc})
2676: and~\eref{e.fullpk} with $k=4$,
2677: this gives when $s<8\,R/9$
2678: \begin{equation}\label{e.sumI}
2679: I_H(A_R)\le O(1)\,\Ps{A_R}\,s^{-5/4+o(1)}.
2680: \end{equation}
2681: where this $O(1)$ factor (as well as those appearing
2682: below) may depend on $k$ and $\theta$.
2683: Since the number of hexagons in $C_\theta$
2684: satisfying $s=s(H)<\rho$ is $O(\rho^2)$,
2685: an easy calculation yields
2686: $$
2687: \sum_{H:s(H)< 8R/9} I_H(A_R)\le O(1)\,\Ps{A_R}\,R^{3/4+o(1)}\,.
2688: $$
2689: 
2690: Now suppose that $H$ is a hexagon satisfying
2691: $s(H)\ge 8\,R/9$.
2692: For $H$ to be pivotal, it is necessary that there would
2693: be $k$ (alternating, if $k>1$) crossings in $C_\theta$
2694: between $\{|z|=10\,k\}$ and $\{|z|=R/2\}$,
2695: there should be $4$ alternating crossings
2696: between $H$ and distance $\dist(H,\p D_R)/2$ from $H$,
2697: and there should be $3$ alternating
2698: crossings
2699: between distance $2\,\dist(H,\p D_R)$ and
2700: distance $R/2$ from a point on $\p D_R$ closest to $H$.
2701: The latter event is governed by the $3$-arm half
2702: plane exponent, whose asymptotic behaviour is described by~\eref{e.halfpk}.
2703: Since $s+\dist(H,\p D_R)=R+O(1)$, we get
2704: $$
2705: I_H(A_R)\le O(1)\,\Ps{A_R}\,(R-s)^{-5/4+o(1)}\,((R-s)/R)^{2+o(1)}\,.
2706: $$
2707: Since for $b\ge 1$ there are $O(b\,R)$ hexagons at distance $\le b$ from
2708: $\{|z|=R\}$, another easy calculation gives
2709: $$
2710: \sum_{H:s(H)\ge  8R/9} I_H(A_R)\le O(1)\,\Ps{A_R}\,R^{3/4+o(1)}\,.
2711: $$
2712: Together, this yields~\eref{e.pivotal} and the proof is complete.
2713: \QED
2714: 
2715: 
2716: \proofof{Theorem~\ref{th:wedges}}
2717: The lower bound was proved in Section~\ref{sec:wedgeslower},
2718: and so only the upper bound needs to be justified.
2719: The proof proceeds like the proof of the upper bound in Theorem~\ref{th:cones},
2720: except that the influence estimates are slightly different.
2721: 
2722: Let $D_R=\{z\in W_\theta:10\,k\le |z|\le R\}$,
2723: $A_R$ be the $k$-arm event in $W_\theta$
2724: between $\{z:|z|=10\,k\}$ and
2725: $\{z:|z|=R\}$, and $H\subset D_R$ be a hexagon.
2726: Let $s=s(H)=\dist (0,H)$, and let $b=b(H)=\dist (H,\p D_R)$,
2727: where we write $\p D_R$ for the boundary of $D_R$ in $C_\infty$,
2728: i.e., the points on $\p W_\theta$ are included.
2729: For $H$ to be pivotal for $A_R$, it is necessary
2730: that the $k$ arm event holds
2731: between distance $10\,k$ and $s/2$ from $0$ (unless
2732: $s/2\le 10\,k$) as well as between
2733: distances $2\,s$ and $R$ (unless $2\,s\ge R$),
2734: that the alternating $4$-arm
2735: event hold between $H$ and distance $b/2$ away from $H$,
2736: and that the alternating $3$-arm event must hold
2737: between distances $2\,b$ and $s/4$ away from a point
2738: in $\p D_R$ closest to $H$ (unless $2\,b\ge s/4$).
2739: There are $O(b's')$ hexagons $H$ satisfying $b(H)\le b'$
2740: and $s(H)\le s'$.
2741: The rest of the proof proceeds like that of Theorem~\ref{th:cones},
2742: and is left to the reader.
2743: \QED
2744: 
2745: 
2746: \proofof{Theorems~\ref{th.no2clusters},
2747: \ref{th.differentclusters},
2748: \ref{th.halfplane} and
2749: \ref{th.halfplane2}}
2750: At any time at which there are 2 infinite white clusters in the plane,
2751:  we must also have
2752: the 4-arm event occuring (with alternating colors) but by
2753: Theorem~\ref{th:cones} (with $k=4$ and $\theta=2\pi$), there are no such times.
2754: This proves Theorem~\ref{th.no2clusters}.
2755: 
2756: At any time at which there are 2 infinite different colored clusters, we must also 
2757: have the 2-arm event occuring (with different colors) but by
2758: Theorem \ref{th:cones} (with $k=2$ and $\theta=2\pi$), the set of
2759: such times has Hausdorff dimension at most $2/3$.
2760: This proves Theorem~\ref{th.differentclusters}.
2761: The other two theorems are similarly proved.
2762: \QED
2763: 
2764: 
2765: 
2766: 
2767: \section{The square lattice} \label{sec:square}
2768: 
2769: We start this section by proving 
2770: Theorem~\ref{th.no3clustersZ2}.
2771: Afterwards, possible ways in which our arguments for
2772: Theorem~\ref{th:except} may be improved to apply to $\Z^2$ 
2773: as well, will be discussed.
2774: 
2775: In the proof of Theorem~\ref{th.no3clustersZ2} we will
2776: use the fact that the $6$ alternating arms exponent is
2777: larger than $2$, or, more precisely, 
2778: that the probability for $6$ alternating arms
2779: between radii $r$ and $R$ is
2780: bounded above by $O(1)\,(r/R)^{2+\eps}$ for some $\eps >0$.
2781: This is essentially due to~\cite[Lemma 5]{KSZ},
2782: but a proof is also given in the appendix (Corollary~\ref{c.56}).
2783: 
2784: 
2785: \proofof{Theorem~\ref{th.no3clustersZ2}}
2786: For $0<r<R$, let $S(r,R)$ be the event
2787: that there are $3$ different clusters that connect the circles
2788: of radii $r$ and $R$ about $0$.
2789: By the above mentioned bound on the alternating $6$-arm probabilities,
2790: We may choose some fixed $\eps>0$ and
2791: some function $\rho=\rho(r)>r$ such that for static critical bond percolation on $\Z^2$,
2792: for all $r$,
2793: \begin{equation}\label{e.rhochoice}
2794: \Pb{ S(r,\rho)}\le\rho^{-2-\eps}\,.
2795: \end{equation}
2796: Consider some bond $e$, and let $F(e)$ be the event that
2797: $e$ is pivotal for $S(r,\rho)$.
2798: Then $\Pb{F(e)}$ is just the influence $I_e(S(r,\rho))$.
2799: Assume that $\Pb{F(e)}\ne 0$.
2800: Note that the events $F(e)$ and $\{e\text{ is open}\}$ are independent
2801: events. This implies that $\Pb{S(r,\rho)\md F(e)}=1/2$,
2802: which one may write
2803: $ \Pb{F(e)\cap S(r,\rho)}=\Pb{F(e)\cap\neg S(r,\rho)}$.
2804: Since this applies to every bond $e$, we conclude that
2805: the expected number of pivotals on the event $S(r,\rho)$
2806: is half of the total influence $I(S(r,\rho))$.
2807: However the number of pivotals for $S(r,\rho)$ is bounded
2808: by the total number of edges intersecting the
2809: disk of radius $\rho$ about the origin, which is
2810: certainly $O(\rho^2)$. Thus, 
2811: $$
2812: I(S(r,\rho))\le 2\,\Pb{S(r,\rho)}\,O(\rho^2) =O(\rho^{-\eps}).
2813: $$
2814: Consequently, by Theorem~\ref{th.hdgeneral}, for every $r_0>0$ a.s.\ there
2815: are no exceptional times in which
2816: $\bigcap_{r>r_0} S(r,\rho(r))$ holds.
2817: This proves our theorem.
2818: \QED
2819: 
2820: 
2821: 
2822: \begin{remark}
2823: An alternative way to prove the above result is based on using the fact that the 6-arm 
2824: exponent is strictly larger than 2 together with the fact that the number of different
2825: configurations (counting repetitions) that appear in a ball of radius 
2826: $n$ during the time interval $[0,1]$ has a Poisson distribution
2827: with a parameter which is at most $O(1) n^2$.
2828: \end{remark}
2829: 
2830: 
2831: \bigskip
2832: 
2833: As we will briefly explain below, the proof of
2834: Theorem~\ref{th:except} almost works for bond percolation on the square
2835: grid. In fact, there are several alternative routes by which
2836: the result might perhaps be extended to $\Z^2$:
2837: \begin{enumerate}
2838: \item establishing 
2839: \begin{equation}\label{e.exponeed}
2840: \alpha_2(r)\le O(r^{-\eps})\,\alpha(r)^{2}
2841: \end{equation}
2842:  for $\Z^2$ for some fixed $\eps >0$,
2843: \item improving the estimate~\eref{e.noise},
2844: \item proving the existence of an algorithm
2845: (or a witness which would still permit the use of
2846: Theorem~\ref{t.noise}) with smaller revealment,
2847: \item extending Smirnov's theorem to $\Z^2$.
2848: \end{enumerate}
2849: 
2850: Note that the weaker version of~\eref{e.exponeed}
2851: $\alpha_2(r)\le \alpha(r)^2$ follows from either the Harris-FKG 
2852: inequality or Reimer's inequality~\cite{Reimer}.
2853: Kesten and Zhang have proved some related strict inequalities
2854: between exponents~\cite{KZ}, but it seems that their
2855: methods are not sufficient to prove~\eref{e.exponeed}.
2856: 
2857: We now explain why~\eref{e.exponeed} in the $\Z^2$ setting implies
2858: exceptional times for $\Z^2$.
2859: First we want to have the revealment
2860: for the algorithm determining $\croRr$ bounded by
2861: $O(1)\,\alpha_2(r)\,\alpha(r,R)$.
2862: One problem seems to be that
2863: the bound on the revealment
2864: for the triangular grid involves the
2865: summand featuring $\alpha^+$, which is relatively negligible, while
2866: on $\Z^2$, we do not know how to prove that the other summand
2867: dominates. The fix is to replace the deterministic $R$
2868: by a random $R_t'\in[R,2R]$.
2869: The random variable $R_t'$ will depend on some extra random bits,
2870: that we add, and these random bits also evolve in time.
2871: We construct the dependence of $R'$ on these bits
2872: so that $R'$ can be calculated by an algorithm with very
2873: small revealment.
2874: This is rather easy to arrange, because we are not limited 
2875: in the number of bits that we may take. 
2876: If we consider an edge whose distance from the origin $a$ is
2877: in the range $[R/2,2R]$, then the probability
2878: that the edge is examined given $R'$ is at most
2879: $O(1)\,\alpha_2(R'-a)\,1_{R'\ge a-1}$.
2880: By~\eref{e.qm}, this is at most
2881: $O(1)\,\alpha_2(R)\,\alpha_2(R'-a,R)^{-1}\,1_{R'\ge a-1}$.
2882: The probability that $|R'-a|\le 2^j$ is at most $O(1)\,2^{j}/R$.
2883: We also know that $\alpha_2(r_1,r_2)^{-1}\le O(1)\,(r_2/r_1)^{1-\eps'}$
2884: for some $\eps'>0$, by Reimer's inequality~\cite{Reimer} and~\eref{e.fivearm}.
2885: It follows that the probability that such an edge is examined is 
2886: $O(1)\,\alpha_2(R)$.
2887: The $r^{o(1)}$ factor in Theorem~\ref{t.annulus}
2888: is easily replaced by an $O(1)$ factor, if we use Proposition~\ref{p.qm}
2889: in the course of the proof.
2890: Then we get~\eref{e.zz} for the square grid, but without the $r^{o(1)}$ factor.
2891: We may then choose the dependence
2892: between $r$ and $t$ such that
2893: $\alpha(r)\approx t\,r^{\eps/2}$, where $\eps$ is the constant
2894: in~\eref{e.exponeed}. The rest is immediate from~\eref{e.zz}, since
2895: clearly $r^{-\eps/2}\le O(1)\,t^{\eps'}$ for some $\eps'>0$.
2896: 
2897: A consequence of this argument, which applies without assuming~\eref{e.exponeed},
2898: is that for bond percolation on $\Z^2$ we have
2899: $$
2900: \Pb{V_{t,R}\cap V_{0,R}} \le O(t^{-1})\,\Pb{V_{t,R}}^2.
2901: $$
2902: This gives yet another illustration as to how close the result for
2903: $\Z^2$ seems to be --- if the $t^{-1}$ term was improved to
2904: $t^{-1+\eps}$, that would have been enough.
2905: Consequently, significant improvements in the algorithm or in~\eref{e.noise}
2906: would also be sufficient.
2907: 
2908: 
2909: 
2910: 
2911: \section{Some open questions} \label{sec:questions}
2912: 
2913: Following are a few questions and open problems suggested by the
2914: present paper.
2915: 
2916: \begin{enumerate}
2917: \item For the results in Theorems \ref{th:wedges} and \ref{th:cones},
2918: what is the Hausdorff dimension of the set of exceptional times 
2919: in question? We tend to believe that the answer is the upper bound.
2920: In particular, is the upper bound of $31/36$ in
2921: Theorem \ref{th.hd} the correct answer?
2922: \item Prove that there exist exceptional times for percolation
2923: on the square lattice. (See Section~\ref{sec:square} for a discussion.)
2924: \item What is the best $\gamma$ for which Theorems~\ref{th:crossingquantsquare}
2925:  and~\ref{th:crossingquant} hold?
2926: \item
2927: What is the best revealment of an algorithm determining
2928: the event $\cro_R$ in Theorem~\ref{t.square}?
2929: \item
2930: What is the sharp form of Theorem~\ref{t.noise}?
2931: \item
2932: What are the properties of the infinite cluster at an exceptional
2933: time at which it exists? 
2934: For example, what is the growth rate of the number of vertices
2935: in the Euclidean disk of radius $r$ around the origin which
2936: belong to the cluster of the origin at the first time
2937: $t\ge 0$ in which the cluster is infinite?
2938: Is the growth rate the same at all exceptional times?
2939: \item
2940: What is the relationship between the exceptional infinite
2941: cluster and the incipient infinite cluster?
2942: \end{enumerate}
2943: 
2944: \appendix
2945: \section{Appendix: Quasi-multiplicativity}\label{s.appendix}
2946: 
2947: In this appendix, we discuss the $k$-arm probabilities
2948: and prove that they satisfy the corresponding analogue of 
2949: the relation~\eref{e.multip}.
2950: For $R>0$, let $H_R$ be the union  of the hexagons intersecting $B(0,R)$.
2951: For $R>r>0$ let $A_j(r,R)$ denote the event that
2952: there are at least $j$ crossings from $\p H_r$ to $\p H_R$, of alternating colors.
2953: The following result refers to critical site percolation on the triangular
2954: grid and critical bond percolation on the square grid.
2955: 
2956: \begin{proposition} \label{p.qm}
2957: Let $j>0$ be even. There is a constant $C$, depending only on $j$,
2958: such that for all $r<r'<r''$
2959: \begin{equation}\label{e.qm}
2960: C^{-1}\, \Pb{A_j(r,r'')}\le
2961: \Pb{A_j(r,r')}\,\Pb{A_j(r',r'')}\le C\, \Pb{A_j(r,r'')},
2962: \end{equation}
2963: and $\Pb{A_j(r,2\,r)}>1/C$ if $\Pb{A_j(r,2\,r)}>0$
2964: (i.e., if $r$ is large enough to allow $j$ crossings).
2965: Moreover, a corresponding statement holds for critical bond percolation on the
2966: square grid which alternate between primal and dual crossings.
2967: \end{proposition}
2968: 
2969: This theorem would have been a useful tool in~\cite{SW},
2970: had it been available. Instead, the authors of that paper 
2971: proved a weaker form of this which was good enough for their
2972: purposes. Our proof below uses techniques from~\cite{K2},\cite{LSWup2}
2973: and \cite{SW}.
2974: Indeed, the entire results of the appendix do follow from the ideas
2975: of~\cite{K2}. We include them here for completeness, and for ease
2976: of reference. Additionally, though the basic ideas are the same,
2977: in several respects our treatment is a bit different from~\cite{K2}.
2978: 
2979: Below, we will work in the setting of the triangular grid.
2980: The proof for the square grid is essentially the same.
2981: In the setting of the triangular grid, there is the color exchange trick~\cite{KSZ,ADA},
2982: which shows that the probability for having alternating crossings is comparable to the
2983: probability of any color sequence as long as both colors are present.
2984: In the setting of the square grid, as far as we know, such a trick does not exist.
2985: At the end of the appendix we will explain how the proof of Proposition~\ref{p.qm}
2986: can be generalized to any color sequence.
2987: 
2988: In the following, an interface
2989: from $\p H_r$ to $\p H_R$ is an oriented simple path in the
2990: hexagonal grid that has one color of hexagons adjacent to it on the
2991: right, and the opposite color adjacent to it on the left.
2992: Thus, it is the common boundary of a black crossing and a white crossing.
2993: 
2994: 
2995: 
2996: For $R>r>1$, consider all the interfaces crossing from $\p H_{r}$ to
2997: $\p H_{R}$, and define $s(r,R)$ to be the least distance between any
2998: pair of endpoints of two interfaces on $\p H_R$.
2999: If there are no interfaces, we take $s(r,R)=\infty$.
3000: Note that $s(r,R)$ is monotone non-increasing in $r$.
3001: This quantity will roughly measure the \lq\lq quality\rq\rq\
3002: of the interfaces; when $s(r,R)$ is comparable to $R$, the interfaces
3003: are well separated, and, as we will see, easier to extend.
3004: 
3005: \begin{lemma}\label{l.apart}
3006: For all $a\in(0,1)$, $R>0$, $\delta>0$
3007: $$
3008: \Pb{s(a\,R,R)<\delta\,R}\le C\,\delta^\eps,
3009: $$
3010: where $C=C(a)$ is a constant depending only on
3011: $a$ and $\eps>0$ is an absolute constant.
3012: \end{lemma}
3013: 
3014: The lemma probably follows from \cite[Lemma 2]{K2},
3015: but since the proof is rather short, we include a proof
3016: for completeness.
3017: 
3018: \proof
3019: We prove this in the case $a=1/2$.
3020: The general case is essentially the same.
3021: Let $\alpha\subset\p H_R$ be an arc of diameter 
3022: $R/3$, and let $Y$ be the set of points
3023: in $H_R$ at distance at most $R/3$ from
3024: $\alpha$.
3025: Let $\alpha_1$ be one of the two
3026: arcs in $\p Y\cap\p H_R\setminus\alpha$.
3027: Let $\beta_1,\beta_2,\dots,\beta_k$
3028: be the interfaces crossing
3029: from $\p Y\setminus\p H_R$ to $\alpha$,
3030: ordered so that for $i_1<i_2\le k$,
3031: the interface $\beta_{i_1}$ separates
3032: $\alpha_1$ from $\beta_{i_2}$ in $\closure Y$.
3033: Fix a positive integer $i$
3034: and condition on $i\le k$ and on $\beta_i$.
3035: Let $\hat\beta_i$ denote the union of the hexagons adjacent
3036: with $\beta_i$.
3037: Then the
3038: percolation in the connected component $Y_i$ of
3039: $Y\setminus\hat \beta_i$ separated from $\alpha_1$
3040: by $\beta_i$ remains unbiased.
3041: Suppose that the hexagons in $\hat\beta_i$
3042: adjacent to $Y_i$ are white, say.
3043: Let $z_i$ denote the endpoint of $\beta_i$ on
3044: $\alpha$.
3045: The RSW theorem implies that
3046: there is some constant $\eps>0$ such that with conditioned
3047: probability $1-O(1)\,\delta^\eps$
3048: there is a white crossing 
3049: in $Y_i\setminus B(z_i,\delta\,R)$ from
3050: $\p\hat\beta_i$ to $\p H_R$.
3051: On that event, it is clear that
3052: if $k\ge i+1$, then $|z_i-z_{i+1}|\ge \delta\,R$.
3053: We conclude that
3054: $$
3055: \PB{k\ge i+1,\,|z_i-z_{i+1}|\le\delta\,R\md
3056: k\ge i,\,\beta_i}\le O(1)\,\delta^\eps.
3057: $$
3058: The RSW theorem also implies that there is conditioned probability
3059: bounded away from zero that $k=i$
3060: given $k\ge i$ and $\beta_i$ (this would be guaranteed by
3061: an appropriate crossing in $Y_i$ from $\p\hat\beta_i$
3062: to $\p H_R\setminus\alpha$). Therefore,
3063: $\Pb{k\ge i}\le c^i$ for some constant $c<1$.
3064: Thus, 
3065: \begin{multline*}
3066: \PB{k\ge i+1,\,|z_i-z_{i+1}|\le\delta\,R}
3067: \\
3068: =
3069: \PB{k\ge i+1,\,|z_i-z_{i+1}|\le\delta\,R\md k\ge i}\,
3070: \Pb{k\ge i}= O(1)\,c^i\,\delta^\eps\,.
3071: \end{multline*}
3072: We sum this over all $i=1,2,\dots$, and over an appropriate
3073: covering of $\p H_R$ by $O(1)$ arcs $\alpha$ of diameter $R/3$.
3074: %The lemma follows using the monotonicity of $s(r,R)$ in $r$.
3075: The lemma follows.
3076: \QED
3077: 
3078: 
3079: Next, we prove a statement
3080: saying, roughly, that if the crossings are \lq\lq reasonably good\rq\rq,
3081: then there is a conditioned probability bounded away
3082: from zero that they extend and
3083: the extensions are \lq\lq very good\rq\rq.
3084: 
3085: \begin{lemma}\label{l.improve}
3086: For every $j>0$ even, there is a constant $\bar\delta=\bar\delta(j)>0$
3087: such that 
3088: for every $\delta>0$ there is some constant $c(\delta)>0$, depending only on $\delta$,
3089: such that when $R>r$,
3090: $$%\begin{equation}\label{e.cdel}
3091: \PB{A_j(r,4\,R)\cap\{s(r,4R)>4\,\bar\delta\,R\}\md  A_j(r,R)\cap\{ s(r,R)>\delta\,R\} } > c(\delta)\,.
3092: $$%\end{equation}
3093: \end{lemma}
3094: 
3095: \proof
3096: Set $S=\closure{H_R\setminus H_r}$.
3097: We assume that $A_j(r,R)$ holds and
3098: that $s(r,R)>\delta\,R$.
3099: Let $\gamma_0,\dots,\gamma_{k-1}$  ($k\ge j$) be the collection of all interfaces crossing
3100: from $\p H_r$ to $\p H_R$, in counterclockwise order,
3101: where we choose the indexes so that $\gamma_0$ has white hexagons
3102: on the right hand side. 
3103: (The interfaces are oriented from $\p H_r$ to $\p H_R$.)
3104: In the following, we set $\gamma_i:=\gamma_{i'}$
3105: when $i\notin \{0,1,\dots,k-1\}$ and $i'=i\mod k$.
3106: Set $\Gamma=\bigcup_{i\in \N}\gamma_i$.
3107: How does conditioning on the interfaces $\gamma_0,\dots,\gamma_{k-1}$
3108: affect the percolation process? 
3109: Note that the fact that there are no more than $k$
3110: interfaces means that for each $i\in\N$ there is a white crossing
3111: in $S\setminus\Gamma$
3112: from the right hand side of $\gamma_{2i}$ to
3113: the left side of $\gamma_{2i-1}$ and 
3114: a black crossing in $S\setminus\Gamma$ from the left side of
3115: $\gamma_{2i}$ to the right side of $\gamma_{2i+1}$.
3116: Otherwise, the configuration is unbiased on hexagons
3117: that are not adjacent to these interfaces.
3118: 
3119: Let $z_i$ be the endpoint of $\gamma_i$ on $\p H_R$, $i\in\N$.
3120: For $i=0,1,\dots,{j-1}$, let $w_i$ be a point $\p H_R$ that
3121: is roughly in the center of the counterclockwise arc from
3122: $z_i$ to $z_{i+1}$. Then $|w_i-z_i|\ge\delta\,R/5$,
3123:  and the same is true for $|w_i-z_{i+1}|$.
3124: It is easy to see that there exist disjoint simple paths
3125: $\beta_0,\dots,\beta_{j-1}$ satisfying the following.
3126: (See Figure~\ref{f.betas}.)
3127: (1) Each $\beta_i$ is a path in $\closure {H_{4R}\setminus H_R}$ 
3128: from $w_i$ to a point $w_i'\in\p H_{4R}$.
3129: (2) The points $w_i'$ are roughly equally spaced on
3130: $\p H_{4R}$.
3131: (3) Each $\beta_i\cap H_{2R}$ is contained in the line
3132: through the origin containing $w_i$,
3133: and each $\beta_i\setminus H_{3R}$ is contained
3134: in the line through the origin containing $w_i'$.
3135: (4) The distance from each of these paths to any other
3136: path is at least $c_1\,\delta\,R$,
3137: where $c_1\in(0,1/5)$ is some universal constant.
3138: (5) Each $\beta_i$ has length at most constant
3139:  times $R$, where the constant may depend on $j$.
3140: For each $i\in\{0,1,\dots,j-1\}$ let
3141: $\alpha_i$ be the arc of a circle whose center
3142: is $z_i$, that has $w_i$ as endpoint,
3143: that has the other endpoint on $\gamma_i\cup\gamma_{i+1}$,
3144: that is otherwise disjoint from $\gamma_i\cup\gamma_{i+1}$
3145: and is contained in $S$.
3146: 
3147: \begin{figure}
3148: \SetLabels
3149: \E(.5*.5)$H_R$\\
3150: \T(.5*.98)$\p H_{4R}$\\
3151: \endSetLabels
3152: %\ShowGrid
3153: \centerline{\epsfysize=2.5in%
3154: \AffixLabels{%
3155: \epsfbox{betas.eps}%
3156: }%
3157: }
3158: \begin{caption} {\label{f.betas}The paths $\beta_i$.}
3159: \end{caption}
3160: \end{figure}
3161: 
3162: Let $\hat\beta_i$ be the connected component containing $w_i'$ of
3163: the complement of $\Gamma$ in the $c_1\,\delta\,R/20$ neighborhood
3164: of $\beta_i\cup\alpha_i$.
3165: If $i\in\{0,1,\dots,j-1\}$ is odd, let $F_i$ denote the event that
3166: there is a white crossing
3167: in $\hat\beta_i$ from $\Gamma$ to $\p H_{4R}$.
3168: Similarly, if $i$ is even, let $F_i$ denote the event that
3169: there is a black crossing in $\hat\beta_i$ from
3170: $\Gamma$ to $\p H_{4R}$.
3171: It is easy to see that if $\bigcap_{i=0}^{j-1}F_i$ holds,
3172: then $A_j(r,4R)$ holds as well.
3173: The RSW theorem implies that $\Pb{F_i}$ is bounded from
3174: below (depending on $\delta$) for a percolation process that is unbiased.
3175: But, as we have seen,
3176: the percolation on $S$ between $\gamma_i$ and $\gamma_{i+1}$
3177: is only conditioned on having a crossing of the appropriate
3178: color. By the Harris-FKG inequality, this is positively
3179: correlated with $F_i$. By independence on
3180: disjoint sets, the different $F_i$ are independent
3181: given $\Gamma$
3182: (assuming, as we may, that the distance between the different
3183: sets $\hat\beta_i$ is significantly larger than the scale of the lattice). 
3184: We conclude that for some $c(\delta)>0$,
3185: $$
3186: \PB{A_j(r,4\,R)\md  A_j(r,R)\cap\{ s(r,R)>\delta\,R\} } > c(\delta)\,.
3187: $$
3188: 
3189: 
3190: \begin{figure}
3191: \SetLabels
3192: %
3193: \L(.83*.85)$\p H_{4R}$\\
3194: \E(.01*.01)$\hat\beta_i$\\
3195: \L\B(.39*.7)$\p H_{3.5\,R}$\\
3196: \endSetLabels
3197: %\ShowGrid
3198: \centerline{\epsfysize=2.5in%
3199: \AffixLabels{%
3200: \epsfbox{bent.eps}%
3201: }%
3202: }
3203: \begin{caption} {\label{f.bent}A bent strip connecting with a channel.
3204: Indicated are the leftmost crossing of
3205: the intersection of the bent strip and the channel that connects
3206: to $H_r$ and a reasonably likely crossing from it to $\p H_{4R}$ in the bent strip.}
3207: \end{caption}
3208: \end{figure}
3209: 
3210: Taking care of the condition $s(r,4R)\ge 4\,\bar\delta\,R$ is not too hard.
3211: Suppose that in the above we truncate the paths
3212: $\beta_i$ and the neighborhoods $\hat\beta_i$ by intersecting
3213: them with $H_{3.5\,R}$.
3214: We then condition on the \lq\lq leftmost\rq\rq\
3215: crossing in $\hat\beta_i$.
3216: See Figure~\ref{f.bent}. 
3217: All this takes place within $H_{3.5\,R}$.
3218: The conditional probability that these crossings in
3219: the $\hat\beta_i$'s connect to
3220: $\p H_{4\,R}$ is bounded away from
3221: zero by a function of $j$ only (specifically, not $\delta$).
3222: Thus, Lemma~\ref{l.apart} and the monotonicity
3223: of $s(r,R)$ in $r$ shows that if
3224: $\bar\delta=\bar\delta(j)>0$ is chosen small, with conditional
3225: probability at least $1/2$ we are also likely to have $s(r,4R)\ge4\,\bar\delta\,R$,
3226: as required.
3227: \QED
3228: 
3229: 
3230: 
3231: Set 
3232: $$
3233: f(r,R):=\Pb{A_j(r,R)}, 
3234: \qquad
3235: g_\delta(r,R):=\Pb{A_j(r,R)\cap \{s(r,R)>\delta\,R\}}.
3236: $$
3237: 
3238: 
3239: \begin{lemma}\label{l.plenty}
3240: There is a constant $\CCb(j)>0$, depending only on $j$,
3241: such that for $R\ge 4\,r$
3242: $$
3243: \CCb (j)\, g_{\bar \delta}(r,R)\ge f(r,R)\,,
3244: $$
3245: where $\bar\delta=\bar\delta(j)$ is the constant introduced in
3246: Lemma~\ref{l.improve}.
3247: \end{lemma}
3248: 
3249: \proof
3250: We assume that $f(r,R)>0$.
3251: Let $\delta>0$ be small.
3252: Set $N=\log_4(R/r)$ and let $m=m_\delta$ be the largest
3253: integer in the range $0\le m\le N-1$
3254: such that
3255: $ g_\delta(r,4^{-i}R)\le f(r,4^{-i}R)/2 $
3256: holds for every integer $i$ in the range
3257: $0\le i< m$. 
3258: Lemma~\ref{l.apart} and independence
3259: on disjoint sets imply
3260: $$f(r,R)-g_\delta(r,R)\le C\,\delta^\eps\, f(r,R/4),$$
3261: and repeated applications of this inequality give
3262: \begin{equation}\label{e.fR}
3263: f(r,R)\le (2C)^m\,\delta^{\eps\, m} f\bl(r,4^{-m} R\br)\,.
3264: \end{equation}
3265: We claim that if $\delta$ is a sufficiently small positive constant, then
3266: \begin{equation}\label{e.gd}
3267: f(r,4^{-m}R) \le \CCc(j)\,g_\delta(r,4^{-m}R)
3268: \end{equation}
3269: for some constant $\CCc(j)$ depending only on $j$.
3270: If $m\le N-2$, this follows with $\CCc(j)=2$ from the definition of $m$.
3271: If $N-2<m\le N-1$, then RSW easily implies
3272: $f(r,4^{-m}R)\ge \CCa (j)$ for some constant $\CCa(j)>0$,
3273: and $f(r,4^{-m}R)-g_\delta(r,4^{-m}R)\le C\,\delta^\eps$ by Lemma~\ref{l.apart},
3274: which gives~\eref{e.gd}.
3275: %
3276: On the other hand, repeated application of Lemma~\ref{l.improve}
3277: gives
3278: \begin{equation}\label{e.fullcircle}
3279: c(\delta)\,c(\bar\delta)^{m-1}\,g_\delta(r,4^{-m}R)
3280: \le g_{\bar\delta}(r,R) \le f(r,R)
3281: \,.
3282: \end{equation}
3283: On combining this with~\eref{e.fR} and~\eref{e.gd}, one obtains
3284: $$
3285: c(\bar\delta)\,\CCc(j)\,\bl(2\,C\,\delta^\eps/ c(\bar\delta)\br)^m \ge c(\delta)\,.
3286: $$
3287: We choose $\delta$ sufficiently small so that $\delta^\eps< c(\bar\delta)/(4\,C)$.
3288: Then the above shows that $m$ is bounded by a function of $\delta$ and $j$.
3289: The proof is now completed by combining~\eref{e.fR},~\eref{e.gd}
3290: and~\eref{e.fullcircle}.
3291: \QED
3292: 
3293: 
3294: \proofof{Proposition \ref{p.qm}}
3295: The proof will be given only for the triangular grid,
3296: since the proof in the setting of bond percolation on the
3297: square grid is essentially the same.
3298: As we remarked above, when $\Pb{A_j(r,2\,r)}>0$,
3299: the RSW theorem easily gives $\Pb{A_j(r,2\,r)}>1/C(j)$,
3300: for some $C(j)>0$.
3301: 
3302: We now assume that $r'>8\,r$ and $r''>8\,r'$.
3303: Suppose that $A_j(r,r'/2)\cap\{s(r,r'/2)>\bar\delta\,r'/2\}$
3304: holds. We also assume that the corresponding event
3305: occurs between $\p H_{2r'}$ and $\p H_{r''}$,
3306: but now we require that the interfaces be well
3307: separated on the inner boundary $\p H_{2r'}$,
3308: instead of on the outer boundary.
3309: As in the proof of Lemma~\ref{l.improve},
3310: it is not too hard to see that conditioned on these
3311: events there is probability bounded away from zero
3312: (by a function of $j$)
3313: that the crossings between $\p H_r$ and $\p H_{r'/2}$
3314: will hook up nicely with the crossings between
3315: $\p H_{2r'}$ and $\p H_{r''}$.
3316: Basically, we only need to arrange that the channels
3317: $\hat\beta_i$ for the inner crossings will cross the corresponding
3318: channels of the outer crossings.  
3319: The proof of the right hand inequality in~\eref{e.qm}
3320: now follows from Lemma~\ref{l.plenty} and the corresponding
3321: statement for crossings with interfaces well-separated in the
3322: inner boundary, which is proved in the same way.
3323: If $r''\le8\,r'$, then the right hand inequality
3324: in~\eref{e.qm} is a consequence of Lemmas~\ref{l.improve}
3325: and~\ref{l.plenty}. A similar proof applies when
3326: $r'\le 8\, r$.
3327: 
3328: It now remains to prove the left hand inequality in~\eref{e.qm}.
3329: If $r''<2\,r'$, then $\Pb{A_j(r',r'')}$ is bounded away from
3330: zero (if we assume $\Pb{A_j(r,r'')}>0$)
3331: and we are done since $\Pb{A_j(r,r'')}\le \Pb{A_j(r,r')}$.
3332: Otherwise, we argue that
3333: $\Pb{A_j(r,r'')}\le \Pb{A_j(r,r')}\,\Pb{A_j(2\,r',r'')}$,
3334: by independence on disjoint subsets,
3335: and $\Pb{A_j(r',2\,r')}\,\Pb{A_j(2\,r',r'')}\le C\,\Pb{A_j(r',r'')}$,
3336: by the right hand inequality in~\eref{e.qm}.
3337: Since $\Pb{A_j(r',2\,r')}$ is bounded away from zero, the left hand
3338: inequality now follows.
3339: \QED
3340: 
3341: We now generalize Proposition~\ref{p.qm} to arbitrary sequences
3342: of crossings.
3343: 
3344: 
3345: \begin{proposition}\label{p.gencol}
3346: Let $j\ge 1$ be an integer, and fix a color sequence $X\in\{\mathrm{black},\mathrm{white}\}^j$.
3347: The probabilities for the existence of $j$ disjoint crossings
3348: whose colors match this sequence in counterclockwise order
3349: also satisfy the inequalities in Proposition~\ref{p.qm}.
3350: The corresponding statement also holds
3351: in the setting of critical bond percolation on the square grid.
3352: \end{proposition}
3353: 
3354: \proof
3355: We start with the easier case where all the colors in the sequence $X$
3356: are the same, say black.
3357: Suppose that $r'>2\,r$ and $r''>2\,r'$.
3358: Consider the event that (a) there are $j$ disjoint black crossings
3359: from $\p H_r$ to $\p H_{r'}$ and (b) there are $j$ disjoint
3360: black crossings from $\p H_{r'}$ to $\p H_{r''}$
3361: and (c) there are $j$ disjoint black circuits separating
3362: $\p H_{r'/2}$ from $\p H_{r'}$
3363: and (d) there are $j$ disjoint black circuits
3364: separating $\p H_{r'}$ from $\p H_{2\,r'}$
3365: and (e) there are $j$ disjoint crossings from
3366: $\p H_{r'/2}$ to $\p H_{2\,r'}$.
3367: Note that if we choose any one path in each of (a)--(e),
3368: we can extract from the union a crossing from $\p H_r$ to 
3369: $\p H_{r''}$. To see that we actually have
3370: $j$ disjoint crossings from $\p H_r$ to $\p H_{r''}$, note that if
3371: we remove any $j-1$ hexagons, then there is still one path
3372: remaining in each of (a)--(e), and consequently, there is
3373: still a crossing from $\p H_r$ to $\p H_{r''}$.
3374: Thus, Menger's theorem
3375: (see~\cite{Diestel})
3376: % (see \cite{Diestel}, page 50)
3377: implies that when (a)--(e) all hold
3378: there are $j$ disjoint crossings from $\p H_r$ to 
3379: $\p H_{r''}$. By the Harris-FKG inequality, the events (a)--(e)
3380: are all positively correlated. By RSW, events
3381: (c)--(e) have probabilities bounded away from zero
3382: (assuming that (a) has positive probability).
3383: The inequality corresponding to the right hand inequality in~\eref{e.qm}
3384: now follows. The corresponding left hand inequality,
3385: as well as the cases where $r'\le 2\,r$
3386: or $r''\le 2\,r'$ are now proved as in the proof of Proposition~\ref{p.qm}.
3387: 
3388: We now assume that both colors appear in $X$,
3389: and indicate the adaptations necessary in the proof of
3390: Proposition~\ref{p.qm} to generalize to the present setting.
3391: The quantity $s(r,R)$ needs to be defined slightly differently.
3392: In the modified definition of $s(r,R)$, still only 
3393: interfaces between crossings of opposite colors are considered. 
3394: Suppose that $\gamma_1$ and $\gamma_2$ are two adjacent interfaces
3395: from $\p H_r$ to $\p H_R$, and that $Q$ is the component
3396: of $H_R\setminus (H_r\cup\gamma_1\cup\gamma_2)$ between them.
3397: Let $\dist(z,Z;Q)$ denote the infimal length of a path from $z$ to $Z$
3398: in $\closure Q$.
3399: For $s>0$ set $W(s):=\{z\in Q:\dist(z,\p H_R;Q)\le s\}$.
3400: The {\bf margin} between $\gamma_1$ and $\gamma_2$ is
3401: defined as the supremum of the set of 
3402: $s>0$ such that any path connecting
3403: $\gamma_1$ and $\gamma_2$
3404: in $W(s)$ has length at least $s$.
3405: Now $s(r,R)$ is redefined as the smallest margin
3406: among any two consecutive interfaces.
3407: 
3408: Lemma~\ref{l.apart} is still valid with this new definition of $s(r,R)$.
3409: In fact, the only change needed in the proof is that instead of
3410: looking for a white crossing 
3411: in $Y_i\setminus B(z_i,\delta\,R)$ from
3412: $\p\hat\beta_i$ to $\p H_R$, one looks for
3413: a crossing in $Y_i\setminus B(z_i',2\,\delta\,R)$,
3414: where $z_i'$ is the last point on the arc
3415: $\p Y\cap\p H_R$, directed away from $\alpha_1$,
3416: that has distance at most $\delta\,R$
3417: from $\beta_i$.
3418: (Here, we assume that $\delta<1/100$, say.)
3419: 
3420: We now explain how this new definition facilitates the
3421: obvious analogue of Lemma~\ref{l.improve}.
3422: Indeed, suppose that $\gamma_1$ and $\gamma_2$
3423: are two adjacent interfaces, there are
3424: at least $j_1\le j$ disjoint
3425: black crossings in the sector
3426: $Q$ of $H_R\setminus H_r$ between $\gamma_1$ and $\gamma_2$,
3427: and the margin between $\gamma_1$ and $\gamma_2$ is at least $s$.
3428: Let 
3429: \begin{equation*}\begin{aligned}
3430: Q_i& :=\{z\in W(s): 2\,i\,s/(2\,j_1)\le \dist(z,\p H_R;Q)\le (2\,i+1)\,s/(2\,j_1)\}
3431: \,,\\
3432: Q_i^* &:=\{z\in W(s): 2\,i\,s/(2\,j_1)\le \dist(z,\gamma_1;Q)\le (2\,i+1)\,s/(2\,j_1)\}
3433: \,,
3434: \end{aligned}\end{equation*}
3435: where $W(s)$ is as above.
3436: We may then consider the event that in each $Q_i$, $i=0,1,\dots,j_1-1$,
3437: there is a black crossing from $\gamma_1$ to $\gamma_2$,
3438: and in each $Q_i^*$, $i=0,1,\dots,j_1-1$
3439: there is a black crossing from $\p H_R$ to 
3440: $\p W(s)\setminus (\gamma_1\cup\gamma_2\cup \p H_R)$,
3441: and moreover, the latter crossings continue through well directed
3442: channels all the way to $\p H_{4R}$, as in the proof of
3443: Lemma~\ref{l.improve}.
3444: An application of Menger's theorem,
3445: as in the monochromatic case above, will then guarantee
3446: that at the end $j_1$ disjoint black crossings between $\gamma_1$ and
3447: $\gamma_2$ will extend all the way to 
3448: $H_{4R}$. A compatible construction is applied to
3449: each of the other pairs of adjacent interfaces.
3450: 
3451: Of course, when we condition on the interfaces
3452: $\gamma_0,\gamma_1,\dots,\gamma_{k-1}$, we do not
3453: know how many crossings we will have between each pair of
3454: adjacent interfaces.
3455:  But $k=O(R/s(r,R))$, and so the number of distinct
3456: patterns in which crossings with color sequence type $X$
3457: can occur is bounded by a function of $\delta$ and $j$.
3458: (Specifically, a pattern for $X$ is a specification of
3459: how many crossings are selected between each pair of adjacent interfaces
3460: to make up the sequence of crossings compatible with $X$.
3461: There may very well be additional crossings that are ignored.)
3462: Thus, the most likely pattern given the interfaces
3463: occurs with probability bounded away from zero given that
3464: there are crossings of color sequence $X$
3465: and the construction may be based on this most likely pattern.
3466: The occurrence of this pattern given the interfaces and
3467: the information about the color of hexagons adjacent to
3468: the right hand sides of the interfaces will be a monotone
3469: function in the collection of white hexagons in the regions
3470: between interfaces that have white hexagons on their boundary,
3471: and monotone in black hexagons in the other regions.
3472: Thus, again, the Harris-FKG inequality is applicable.
3473: (We do not want to condition on the exact number
3474: of crossings between two adjacent interfaces, as this is not
3475: a monotone function of the configuration.)
3476: 
3477: Similarly, when we attempt to glue crossings between two
3478: different annuli, we condition on the interfaces, and then
3479: aim for the most likely pattern in each annulus.
3480: These are essentially the only modifications needed in the proof.
3481: \QED
3482: 
3483: \begin{remark}\label{r.qmc}
3484: The analogous statements for critical percolation in cones and
3485: wedges also holds, with similar proofs. The wedge case is, in
3486: fact, easier.
3487: \end{remark}
3488: 
3489: \begin{remark}\label{r.primus}
3490: It is also clear that the above proof shows that if we prescribe
3491: $j$ specific disjoint arcs on $\p H_R$ and require the crossings
3492: from $\p H_r$ to land on
3493: these arcs, with a prescribed color for every arc,
3494: the probability for this event is at least a positive constant
3495: times the probability to have $j$ crossings with this
3496: sequential color pattern, where the constant only depends
3497: on the smallest angle at $0$ subtended by any of the $j$ arcs
3498: (provided that $R$ is not too small, so that each of the arcs
3499: has at least one hexagon unshared with any other arc, say).
3500: \end{remark}
3501: 
3502: As a further application, we prove the following
3503: result about $5$-arm and $6$-arm exponents in $\Z^2$.
3504: 
3505: \begin{corollary}\label{c.56}
3506: Consider critical bond percolation on $\Z^2$.
3507: For $R>r\ge 1$ let $F(r,R)$ denote the event that 
3508: there are five open crossings between distances $r$ and $R$
3509: from $0$ of types primal, primal, dual, primal, dual,
3510: in circular order.
3511: Then
3512: \begin{equation}\label{e.fivearm}
3513:  C^{-1}(r/R)^{2}\le
3514: \Pb{F(r,R)} \le C\,(r/R)^{2},
3515: \end{equation}
3516: where $C>0$ is a universal constant.
3517: Moreover, the probability that there are $6$ alternating
3518: crossings: primal, dual, primal, dual, primal, dual,
3519: between distances $r$ and $R$ is at most
3520: $C\,(r/R)^{2+\eps}$, for some constant $\eps>0$.
3521: The same statement applies to any sequence obtained
3522: by inserting one additional primal or dual
3523: entry to the list (primal, primal, dual, primal, dual).
3524: \end{corollary}
3525: 
3526: This result is essentially due to~\cite[Lemma 5]{KSZ} (in
3527: the context of site percolation on the triangular grid,
3528: though the proof is equally applicable to $\Z^2$).
3529: They omit some of the details,
3530: because the proof is long and the argument is similar to
3531: the proof of~\cite[Lemma 4]{K2}.
3532: Now, we can easily present an essentially complete
3533: and relatively short argument.
3534: 
3535: \proof
3536: We begin with the basic argument from~\cite{KSZ}.
3537: Divide the boundary of the circle $\p B(0,R)$ into
3538: $5$ equal arcs, $A_1,\dots,A_5$, in counterclockwise order.
3539: For concreteness, let's take each $A_j$ as the arc between
3540: angles $(2\,j-1)\,\pi/5$ and $(2\,j+1)\,\pi/5$.
3541: For a vertex $v\in B(0,R)$,
3542: let $F_v$ be the event that
3543: there are primal (open) crossings from $v$ to $A_1,A_3$ and $A_4$ and
3544: dual crossings from dual vertices adjacent to $v$ to $A_2$ and to $A_5$, 
3545: and the primal crossings are disjoint, except at $v$.
3546: (By planarity, it follows that the dual crossings are disjoint.)
3547: We claim that $F_v$ can happen for at most one vertex in
3548: $B(0,R/2)$. Indeed, suppose that $F_v\cap F_u$ holds,
3549: where $v,u$ are vertices in $B(0,R/2)$.
3550: Let $\alpha_i$, $i=1,3,4$ denote some simple primal crossings from $v$
3551: to the arcs $A_i$, which are disjoint, except at $v$.
3552: Similarly, let $\alpha_i'$, $i=1,3,4$, be the corresponding
3553: paths for $u$.
3554: Since $\alpha_1\cup\alpha_3$ separates
3555: $A_2$ from $A_5$ in $B(0,R)$, it is clear that
3556: $u\in \alpha_1\cup\alpha_3$. Similarly,
3557: $u\in\alpha_1\cup\alpha_4$. Since $\alpha_3\cap\alpha_4=\{v\}\subset\alpha_1$,
3558: it follows that $u\in \alpha_1$. Let $\beta_1$ be the arc of $\alpha_1$
3559: from $u$ to $A_1$. Since  $\beta_1\cup\alpha_3'$ is a path from $A_1$ to $A_3$,
3560: it contains $v$ or separates $v$ from $A_2$ or from $A_5$.
3561: The latter two possibilities are ruled out by the dual crossings
3562: to $A_2$ and $A_5$ starting at dual vertices adjacent to $v$.
3563: Thus, $v\in\beta_1\cup\alpha_3'$, and similarly,
3564: $v\in \beta_1\cup\alpha_4'$. But since $\alpha_3'\cap\alpha_4'=\{u\}\subset\beta_1$,
3565: we conclude that $v\in \beta_1$, which implies $u=v$.
3566: 
3567: We now claim that $F:=\bigcup_{v\in B(0,R/2)} F_v$ has probability bounded away from $0$. 
3568: Consider the event that there is a crossing from $A_1$ to $A_4$, and consider the rightmost
3569: such crossing $\ell$ (in the sense that it separates any other crossing from $A_5$).
3570: If there is an open path from $A_3$ to $\ell$, but there is no open path from $A_3$
3571: to $A_1$ disjoint from $\ell$, then $F_w$ will hold, where
3572: $w$ is the first vertex $v$ along $\ell$ (when $\ell$ goes from $A_1$ to $A_4$)
3573: that connects to $A_3$ in the complement of $\ell$.
3574: Thus, we need to show that there is probability bounded away from zero that this
3575: happens with $w\in B(0,R/2)$.
3576: Let $L_1$ be the line passing through the origin and the midpoint of $A_1$.
3577: Let $L_2$ and $L_3$ be lines parallel with $L_1$ at distance $R/20$ and $R/10$
3578: from $L_1$, on the side of $L_1$ that contains $A_5$.
3579: By RSW, there is probability bounded away from zero for the existence of
3580: a dual-open crossing from $A_1$ to $A_4$ in the strip between $L_2$ and $L_3$.
3581: By conditioning on the leftmost such crossing (the one closest to $L_1$),
3582: it is easy to see that there is probability bounded away from
3583: zero that such a dual crossing exists and is also connected to $A_5$ by a
3584: dual-open path.
3585: If moreover we have a primal crossing from $A_1$ to $A_4$ in the strip
3586: between $L_1$ and $L_2$, which happens with
3587: probability bounded away from zero, then the rightmost primal crossing between 
3588: $A_1$ and $A_4$ will be contained in the strip
3589: between $L_1$ and $L_3$.
3590: Conditioned on the latter event and on the latter crossing $\ell$,
3591: it happens with probability bounded away from zero that
3592: there is a primal crossing from $A_3$ to $\ell$ whose endpoint on $\ell$
3593: (which is its only point on $\ell$) is within distance
3594: $R/5$ of the origin and there is a dual crossing from $A_2$ to a dual
3595: vertex adjacent to $\ell$ that is within distance $R/5$ of the origin.
3596: On that event, $F$ holds. Thus, $\Pb{F}$ is bounded away from zero.
3597: 
3598: It is easy to see that the proof
3599: of Proposition~\ref{p.gencol} implies
3600: that $\Pb{F_v}\le O(1)\,\Pb{F_w}$ for $v,w\in B(0,R/2)$.
3601: Since the events $F_v$ are disjoint, and since their sum is of order $1$,
3602: it follows that each $F_v$, $v\in B(0,R/2)$ has probability of order $R^{-2}$.
3603: In particular $R^2\,\Pb{F_0}$ is bounded away from zero and $\infty$.
3604: Now~\eref{e.fivearm} follows from Remark~\ref{r.primus}.
3605: 
3606: The statements regarding the $6$-arm exponent now follow
3607: from Reimer's inequality~\cite{Reimer}. Alternatively, we may also deduce them
3608: from Remark~\ref{r.primus}, as follows. If we fix arcs 
3609: $A_1,\dots,A_6$ in counterclockwise order on the radius $R$ circle, 
3610: where the crossings are required to land,
3611: and we require a primal crossing to $A_1$ and a dual crossing to $A_2$,
3612: then we may condition on the most counterclockwise primal crossing $\gamma$ connecting to $A_1$,
3613: then sequentially on the most clockwise crossings to $A_2,A_3,\dots,A_5$  of the required type.
3614: The conditional probability for having yet another crossing to $A_6$ is 
3615: still bounded by $O(1)\,(r/R)^\eps$, for
3616: some constant $\eps>0$. Now we may apply Remark~\ref{r.primus}, to complete the proof.
3617: \QED
3618: 
3619: \bigskip
3620: \noindent{\bf Acknowledgments}.
3621: We thank Harry Kesten for useful advice.
3622: 
3623: \begin{thebibliography}{AAA}
3624: \addcontentsline{toc}{section}{\refname}
3625: %%\bibliographystyle{halpha}
3626: 
3627: \bibitem{ADA}
3628: Aizenman, M., Duplantier, B. and Aharony, A.
3629: Path crossing exponents and the external perimeter in
3630: 2D percolation. 
3631: {\em Phys. Rev. Let.}  {\bf 83}, (1999), 1359--1362.
3632: 
3633: \bibitem{BKS}
3634: Benjamini, I., Kalai, G. and Schramm, O.
3635: Noise sensitivity of Boolean functions and applications to percolation.  
3636: {\em Inst. Hautes \'Etudes Sci. Publ. Math.}  {\bf 90}, (1999), 5--43.
3637: 
3638: \bibitem{BeSc}
3639: Benjamini, I. and Schramm, O.
3640: Exceptional planes of percolation.  
3641: {\em Probab. Theory Related Fields}, {\bf 111}, (1998), 551--564.
3642: 
3643: \bibitem{BSW}
3644: Benjamini, I., Schramm, O. and Wilson, D. B.
3645: Balanced Boolean functions that can be evaluated so that
3646: every input bit is unlikely to be read.  
3647: {\em STOC '05: Proceedings of the thirty-seventh annual ACM symposium on 
3648: Theory of computing}, (2005), 244--250.
3649: 
3650: \bibitem{BMW}
3651: Berg, J. van den, Meester, R. and White, D. G. Dynamic Boolean
3652: models. {\em Stochastic Process. Appl.}  {\bf 69}, (1997), 247--257.
3653: 
3654: \bibitem{BrSt} Broman, E. I. and Steif J. E.
3655: Dynamical Stability of Percolation for Some Interacting Particle Systems 
3656: and $\epsilon$-Movability, {\em Ann. Probab.},  {\bf 34}, (2006), 539--576.
3657: 
3658: \bibitem{CN} Camia, F. and Newman, C. M.
3659: Two-dimensional critical percolation: the full scaling limit,
3660: {\em Comm. Math. Phys.},  {\bf 268}, (2006), 1--38.
3661: 
3662: \bibitem{Diestel} Diestel, R.  {\em Graph theory.} Springer-Verlag, 
3663: (1997), New York.
3664: 
3665: \bibitem{Eva89}
3666: Evans, S. N.  Local properties of Levy processes on a totally disconnected group,
3667: {\em J. Theoret. Probab.} {\bf 2} (1989), 209--259. 
3668: 
3669: \bibitem{FG} 
3670: Fitzsimmons P.J. and Getoor R.K. 
3671: On the potential theory of symmetric Markov processes.
3672: {\em Math.\ Ann.\ } {\bf 281}, (1988), 495--512.
3673: 
3674: \bibitem{GS} Getoor, R.K. and Sharp, M.J. 
3675: Naturality, standardness, and weak duality for Markov processes.
3676: {\em Zeits.\  Wahr.\ verw.\ Gebiete} {\bf 67}, (1984), 1--62.
3677: 
3678: \bibitem{Grimmett} Grimmett, G.  {\em Percolation.} Second edition,
3679: Springer-Verlag, (1999), New York.
3680: 
3681: \bibitem{HPS} H\"aggstr\"om, O., Peres, Y. and Steif, J. Dynamical
3682: percolation. {\em Ann.\ Inst.\ Henri Poincar\`e, Probab.\ et Stat.}
3683: {\bf 33}, (1997), 497--528. 
3684: 
3685: \bibitem{HS} Hara, T. and Slade, G. Mean field behavior and the lace 
3686: expansion, in {\em Probability Theory and Phase Transitions}
3687: (ed.~G.~Grimmett), 
3688: Proceedings of the NATO ASI meeting in Cambridge 1993, 
3689: (1994), Kluwer.
3690: 
3691: \bibitem{Harris} Harris, T. E.  A lower bound on the critical 
3692: probability in a certain percolation process. {\em Proc. Cambridge Phil.\ Soc.}
3693: {\bf 56}, (1960), 13--20.
3694: 
3695: \bibitem{Haw84}
3696: Hawkes, J. {\em Stochastic analysis and applications (Swansea, 1983)}, 130--154,
3697: Lecture Notes in Math., 1095, Springer, Berlin, 1984.
3698: 
3699: \bibitem{Kahane}
3700: Kahane, J. P. {\em Some random series of functions.} 
3701: Second edition, Cambridge University Press: 
3702: (1985), Cambridge.
3703: 
3704: \bibitem{Kesten} Kesten, H. {\em Percolation theory
3705: for mathematicians,} Birkhauser, (1982), New York.
3706: 
3707: \bibitem{K2} Kesten, H. Scaling relations for 2D-percolation.
3708: {\em Commun.\ Math.\ Phys.} {\bf 109}, (1987), 109--156.  
3709: 
3710: \bibitem{KSZ}
3711: Kesten, H., Sidoravicius, V. and Zhang, Y. 
3712: Almost all words are seen in critical site percolation on the 
3713: triangular lattice.  
3714: {\em Electron.\ J. Probab.}  {\bf 3}, (1998), no. 10, 75 pp. (electronic). 
3715: 
3716: \bibitem{KZ} Kesten, H. and Zhang, Y. 
3717: Strict inequalites for some critical exponents in 2D-percolation.
3718: {\em J. Statist.\ Phys.} {\bf 46}, (1987), 1031--1055.
3719: 
3720: \bibitem{Lawler} Lawler, G. F. 
3721: {\em Conformally invariant processes in the plane.} 
3722: Mathematical Surveys and Monographs, {\bf 114}. 
3723: American Mathematical Society, (2005), Providence, RI.
3724: 
3725: \bibitem{LSW}
3726: Lawler, G., Schramm, O. and Werner, W. 
3727: One-arm exponent for critical 2D percolation  
3728: {\em Electron.\ J. Probab.}  {\bf 7}, (2002), no. 2, 13 pp. (electronic).
3729: 
3730: \bibitem {LSWup2}
3731: Lawler, G., Schramm, O. and Werner, W. 
3732: Sharp estimates for Brownian non-intersection probabilities.
3733: {\em In and out of equilibrium} (Mambucaba, 2000), 113--131,
3734: Progr. Probab., 51, (2002), Birkh\"auser.
3735: 
3736: \bibitem {PSSW}
3737: Peres, Y., Schramm, O., Sheffield, S. and Wilson,D. B.
3738: Random-turn Hex and other selection games,
3739: {\em Amer. Math. Monthly}, May, 2007, to appear.
3740: 
3741: \bibitem{PS} Peres, Y. and Steif, J. E.
3742: The number of infinite clusters in dynamical percolation,
3743: {\em Probab. Theory Related Fields}, {\bf 111}, (1998), 141--165.
3744: 
3745: \bibitem{Reimer} Reimer, D.
3746: Proof of the van den Berg-Kesten conjecture,
3747: {\em Combin. Probab. Comput.},  {\bf 9}, (2000), 27--32.
3748: 
3749: \bibitem{Schramm}
3750: Schramm, O. Scaling limits of loop-erased random walks and uniform spanning trees.  
3751: {\em Israel J. Math.}  {\bf 118}, (2000), 221--288. 
3752: 
3753: \bibitem{Smirnov}
3754: Smirnov, S. 
3755: Critical percolation in the plane: conformal invariance, Cardy's formula, 
3756: scaling limits. {\em C. R. Acad. Sci. Paris S\'er. I Math.}  
3757: {\bf 333},  (2001), 239--244.
3758: 
3759: \bibitem{Smirnov1}
3760: Smirnov, S.
3761: Critical percolation in the plane. I. conformal invariance and Cardy's formula.
3762: II. continuum scaling limit. (long version), preprint.
3763: 
3764: \bibitem{SW}
3765: Smirnov, S. and Werner, W. 
3766: Critical exponents for two-dimensional percolation.  
3767: {\em Math. Res. Lett.}  {\bf 8},  (2001), 729--744
3768: 
3769: \bibitem {Talagrand} Talagrand, M.
3770: Concentration of measure and isoperimetric inequalities in product spaces, 
3771: {\em  Publ. I.H.E.S.}, {\bf 81} (1995), 73--205.
3772: 
3773: 
3774: 
3775: 
3776: \end{thebibliography}
3777: 
3778: \bigskip
3779: \filbreak
3780: \begingroup
3781: {
3782: \small
3783: \parindent=0pt
3784: 
3785: Microsoft Corporation\\
3786: One Microsoft Way\\
3787: Redmond, WA 98052, USA\\ 
3788: {\tt
3789: {http://research.microsoft.com/\string~schramm/}}
3790: 
3791: \bigskip
3792: 
3793: Department of Mathematics \\
3794: Chalmers University of Technology 
3795: and G\"{o}teborg University \\
3796: Gothenburg, Sweden \\
3797: {\tt
3798: {steif@math.chalmers.se} \\
3799: {http://www.math.chalmers.se/$\sim$steif/}}
3800: }
3801: 
3802: 
3803: \filbreak
3804: 
3805: \endgroup
3806: 
3807: 
3808: \end{document}
3809: 
3810: 
3811: 
3812: