1: \documentclass[12pt]{article}
2: \title{Brownian beads}
3: \author{B\'alint Vir\'ag\footnote{Department of Mathematics, MIT, Cambridge, MA
4: 02139, USA. {\sf balint@math.mit.edu}. Research partially
5: supported by NSF grant \#DMS-0206781.}}
6: \date{May 27, 2003}
7: \def\myabstract{}
8: \oddsidemargin 0in \topmargin 0in \headheight 0in \headsep 0in
9: \textheight 9in \textwidth 6.5in
10: \renewcommand{\baselinestretch}{1.3}
11:
12: \usepackage{amsfonts}
13: \usepackage{graphicx}
14: \usepackage{amsmath}
15: \usepackage{amsthm}
16:
17: \newtheorem{theorem}{Theorem}
18: \newtheorem{definition}[theorem]{Definition}
19: \newtheorem{question}[theorem]{Question}
20: \newtheorem{conjecture}[theorem]{Conjecture}
21: \newtheorem{fact}[theorem]{Fact}
22: \newtheorem{remark}[theorem]{Remark}
23: \newtheorem{proposition}[theorem]{Proposition}
24: \newtheorem{lemma}[theorem]{Lemma}
25: \newtheorem{corollary}[theorem]{Corollary}
26: \newenvironment{theorem*}[1]{\par \trivlist
27: \itemindent 0pt \item[\hskip\labelsep\bf Theorem #1]
28: \it\ignorespaces}{\endtrivlist}
29: \newenvironment{eg}{\par \trivlist
30: \itemindent 0pt \item[\hskip\labelsep\bf Example
31: \refstepcounter{theorem} \thetheorem]\ignorespaces}{\endtrivlist}
32: %\newcommand{\qed}{\hfill\mbox{$\framebox(5,5)[]{}$}}
33: \renewenvironment{proof}{\par \trivlist
34: \itemindent\parindent \item[\hskip\labelsep\sc Proof.]
35: \ignorespaces}{\qed\endtrivlist}
36: \newenvironment{proofof}[1]{\par \trivlist
37: \itemindent\parindent \item[\hskip\labelsep\sc Proof of #1.]
38: \ignorespaces}{\qed\endtrivlist}
39: \renewcommand{\theequation}{%\thesection.
40: \arabic{equation}}
41:
42: \newcommand\mnote[1]{}
43: %\newcommand\mnote[1]{\marginpar{\ \\ \small \tt #1}}
44: \newcommand\lb[1]{\label{#1}\mnote{#1}}
45: \newcommand\bel[1]{{\mnote{#1}}\begin{equation}\label{#1}}
46: %\newcommand\bel[1]{\begin{equation}\lb{#1}}
47: \newcommand\ee{\end{equation}}
48:
49: \newcommand\re[1]{(\ref{#1})}
50: \newcommand{\Strip}{\mathbb S}
51: \newcommand{\eps}{\varepsilon}
52: \newcommand{\Var}{\mbox{Var}}
53: \newcommand{\Z}{\mathbb Z}
54: \newcommand{\R}{\mathbb R}
55: \newcommand{\RR}{\mathbb R}
56: \newcommand{\nis}{\hbox{$\cap$\kern-.6em\lower0em\hbox{$/$}}\,}
57: \newcommand{\C}{\mathbb C}
58: \newcommand{\CC}{\mathbb C}
59: \newcommand{\nats}{\mathbb N}
60: \newcommand{\Zd}{{\mathbb Z}^d}
61: \newcommand{\Rd}{{\mathbb R}^d}
62: \newcommand{\HH}{{\mathbb H}}
63: \newcommand{\hull}{\mathcal H}
64: \newcommand{\UU}{\mathbb U}
65: \newcommand{\DD}{\mathbb D}
66: \newcommand{\FF}{\mathcal F}
67: \newcommand{\GG}{\mathcal G}
68: \newcommand{\Ubar}{\overline {\mathbb U}}
69: \newcommand{\Hbar}{\overline {\mathbb H}}
70: \newcommand{\ev}{{\mathbf E}}
71: \newcommand{\pr}{{\mathbf P}}
72: \newcommand{\one}{{\mathbf 1}}
73: \newcommand{\gi}{\,|\,}
74: \newcommand{\as}{\mbox{\hspace{.3cm} a.s.}}
75: \newcommand{\dist}{\mbox{dist}}
76: \newcommand{\Ghat}{{\hat G}}
77: \newcommand{\CBM}{\sf CBM}
78: \newcommand{\BE}{\mathsf{BE}}
79: \newcommand{\bm}{\mathsf{BM}}
80: \newcommand{\BB}{\mathsf{BB}}
81: \newcommand{\nbb}{\mathsf{NBB}}
82: \newcommand{\ic}{IC}
83: \newcommand{\Rnn}{\R_{\ge 0}}
84: \newcommand{\Dbar}{{\overline D}}
85: \newcommand{\cd}{\Rightarrow}
86: \newcommand{\omu}{{\overline \mu}}
87: \newcommand{\onu}{{\overline \nu}}
88: \newcommand{\diam}{\operatorname{diam}}
89: \newcommand{\capz}{\mbox{\rm cap}_0}
90: \newcommand{\capo}{\mbox{\rm cap}_1}
91:
92: \usepackage{natbib}
93: \bibliographystyle{balint}
94:
95: \begin{document}
96: \maketitle
97: %\footnotetext[2]
98: %{{\it AMS} 1991 {\it subject classifications. } Primary ;
99: %secondary .}
100: %\footnotetext[3]{
101: %{\it Key words and phrases.} Brownian motion}
102: \begin{abstract} \noindent
103: We show that the past and future of half-plane Brownian motion
104: at certain cutpoints are independent of each other after a
105: conformal transformation. Like in It\^o's excursion theory, the
106: pieces between cutpoints form a Poisson process with respect to
107: a local time. The size of the path as a function of this local
108: time is a stable subordinator whose index is given by the
109: exponent of the probability that a stretch of the path has no
110: cutpoint. The index is computed and equals $1/2$.
111: \end{abstract}
112:
113: \section{Introduction}
114:
115: The mathematical theory of conformally invariant planar stochastic
116: models has seen great progress in the recent years. The goal of
117: this paper is to consider Brownian motion, the first example of a
118: conformally invariant process, and further explore its conformal
119: structure.
120:
121: Let $B$ denote Brownian motion started at zero and conditioned to
122: stay in the upper half plane; we call this distribution $\BE$, or
123: half-plane excursion. This is a transient process, and he path has
124: cut-points \citep{burdzy89}, that is points which, if removed,
125: make the image of $B$ disconnected. We call the segments of the
126: paths between consecutive cutpoints {\bf Brownian beads}.
127:
128: For a given cutpoint, the complement of the past path has one
129: infinite connected component. This can be mapped back to the half
130: plane via a conformal homeomorphism. It is convenient to normalize
131: this map so that it has derivative 1 at $\infty$ and takes the
132: cutpoint to 0; we call this map the {\bf conformal shift}.
133: %
134: \begin{figure}
135: \label{f2} \centering
136: \includegraphics[height=3in]{bbf1}
137: \caption{Brownian excursion in $\HH$ and its cutpoints}
138: \end{figure}
139: %
140: Our first goal is to show that Brownian beads are independent
141: of each other after a conformal transformation. More precisely,
142: for a cutpoint $g$ let $f$ denote the corresponding conformal
143: shift and let $\beta$ the bead starting at $g$ (if any).
144: \begin{theorem}\lb{ppp}
145: There exists a local time $\lambda$ supported on cutpoints so that
146: $(f_{\lambda}(\beta_{\lambda}), \lambda \ge 0)$ is a Poisson
147: point process.
148: \end{theorem}
149: Here $f_\lambda$ applied to a path maps the image of the path
150: according to the conformal map, and changes time-parameterization
151: according to local Brownian scaling (Section \ref{s.conformal}).
152: The intensity measure of the Poisson point process is a
153: $\sigma$-finite measure on paths conditioned to start without
154: cutpoints; we call this measure the bead process. In Section
155: \ref{s.properties} we prove a Markov property for the bead process
156: and show that it has finite lifetime.
157:
158: Theorem \ref{ppp} may be thought of as a two-dimensional analogy
159: of It\^o's theorem about Brownian excursions. Recall that
160: excursions of 1-dimensional Brownian motion are the segments of
161: the path between two consecutive visits to 0. For a zero $g'$ let
162: $\beta'$ denote the excursion starting at $g'$, and let $f'$
163: denote the map that shifts the future of the path by $-g'$.
164: \begin{theorem}[It\^o]
165: There exists a local time $\lambda'$ on zeros so that
166: $(f'_{\lambda'}(\beta'_{\lambda'}), \lambda' \ge 0)$ is a Poisson
167: point process.
168: \end{theorem}
169: The local time $\lambda'$ at zero has the nice property that the
170: $(g'(\lambda'),\lambda'\ge 0)$ process, or in other words the
171: total duration of excursions up to the zero $g'$ is a stable
172: subordinator of index $1/2$. It turns out that the analogous
173: statement holds in our setting. Let $a(t)$ denote the half-plane
174: capacity of the path up to time $t$; half-plane capacity is a way
175: to measure the size of subsets of $\HH$ (see Section
176: \ref{s.conformal}). Then
177: \begin{theorem}
178: The process $(a\circ g(\lambda), \lambda \ge 0)$ is a stable
179: subordinator of index $\alpha$.
180: \end{theorem}
181: Another property of beads which is shared by It\^o excursions is
182: that the process $(\beta_\lambda,\,\lambda\ge 0)$ determines
183: $(B(t),\,t\ge0)$ (see Remark \ref{determine}).
184:
185: In Section \ref{s.exponent} we show that the index $\alpha$ can be
186: expressed as an exponent.
187:
188: \begin{theorem}\lb{exponent} We have $\pr(\mbox{B has no cuttime in }(1,t))
189: = t^{-\alpha + o(1)}.$
190: \end{theorem}
191:
192: Finally, we compute $\alpha$.
193: \newcommand{\indexhalfthm}
194: {The Brownian bead index $\alpha$ equals $1/2$.}
195: \begin{theorem}\lb{indexhalf}
196: \indexhalfthm
197: \end{theorem}
198: Although the $1/2$ here is equal to the index for It\^o's
199: excursions, it is a coincidence. Neither is it a consequence of
200: Brownian scaling. Unlike the proof of most exponents, this theorem
201: does not rely on SLE processes (nor exponents that use these
202: processes). The proof uses beads, not the exponent representation;
203: I am grateful to W. Werner with whom I discussed ideas of this
204: kind. A related exponent governing the Hausdorff dimension of
205: cuttimes for $\mbox{SLE}_6$ was computed by \cite{beffara}, and it
206: seems possible to derive the value of $\alpha$ directly from his
207: results. \cite{math.PR/0302115} also computed an exponent related
208: to $\alpha$ using SLE$(\kappa,\rho)$ processes (in a paper with
209: beautiful ideas and computations but few rigorous proofs). The
210: advantage of the proof here is that it is more elementary and
211: conceptual; it relies only on the special intersection exponent
212: $\xi(2,1)$, whose value was determined by \cite{lawler95}. His
213: proof covers even the 3-dimensional case.
214:
215: In the proof, we will use the following fact about Brownian
216: excursion. Let $A$ be a compact subset of $\overline \HH\setminus
217: \{0\}$ so that the sets $\RR \cup A$ and $\HH\setminus A$ are
218: connected. Let $f:\HH\setminus A\to\HH$ be a conformal
219: homeomorphism fixing $\infty$ with $f'(\infty)=1$.
220: \begin{proposition}\lb{mainhit}
221: $\quad\pr(B\mbox{ avoids } A) =f'(0)$.
222: \end{proposition}
223:
224: Further interesting properties of Brownian motion have recently
225: been found using the fact that its outer boundary has the same
226: distribution as (the outer boundary of) some SLE processes, see
227: \cite*{lwsrest} and the references therein. Some of methods used
228: in this paper appear independently in the recent literature; in
229: particular, see \cite{dubedat} and \cite{loopsoup}.
230:
231: Many of the arguments of this paper carry over to ordinary
232: Brownian motion, more precisely, Brownian excursion from a
233: boundary point to an interior point of a domain. The analysis
234: there is a bit more difficult since there is no scale invariance;
235: we do not follow this avenue here.
236:
237: In Section \ref{properties} we discuss some simple properties of
238: Brownian excursions in planar domains that we will need later. We
239: also show Proposition \ref{mainhit} there. In Section \ref{s.ctf}
240: we introduce the cuttime filtration and prove a Markov property
241: with respect to this filtration. Section \ref{s.conformal} reviews
242: the facts we need from conformal geometry, and introduces some
243: basic semigroups of paths. In Section \ref{s.independent} we prove
244: the Poisson point process decomposition of Theorem \ref{ppp}, and
245: we define the bead process, a Brownian excursion in the half plane
246: conditioned to start without cutpoints. In the next section we
247: show that this special process does not need infinite time to
248: start off, a fact that is not clear from the definition. In the
249: last two sections, \ref{s.exponent} and \ref{s.index}, we prove
250: Theorems \ref{exponent} and \ref{indexhalf}; we also give some
251: open questions and conjectures.
252:
253: \section{Properties of Brownian excursion}\lb{properties}
254:
255: Let $D$ be a {\bf regular domain}, i.e. a simply connected open
256: subset of the plane whose boundary is locally connected. Let
257: $\partial D$ denote the Caratheodory boundary, i.e. the set of
258: prime ends of $D$.
259:
260: For $\{a\},Z\subset D\cup \partial D$, and $a\not=\infty$, let
261: $\BE(a,Z,D)$ denote Brownian motion started at $a$ and conditioned
262: to hit $Z$ no later than $\partial D$. If this event has positive
263: probability, then this definition is precise; otherwise it can be
264: made precise by considering $h$-processes or by taking a limit. A
265: special case is $\BE(0,\infty,\HH)$, the half-plane excursion,
266: which we will often abbreviate $\BE$. The coordinates of $\BE$ are
267: Brownian motion and an independent 3-dimensional Bessel process.
268:
269: The two most important properties of $B\sim\BE(a,Z,D)$ are
270: restriction and conformal invariance. Restriction says that if
271: $D'\subset D$, then on the event that $B$ stays in $D'$, it
272: has distribution $\BE(a,Z',D')$, where $Z'=Z\cap(D'\cup
273: \partial D')$. Conformal invariance says that if
274: $f:(a,Z,D)\to (a',Z',D')$ is a conformal homeomorphism between two
275: regular domains, then after a time change (\ref{timechange}) the
276: image of $B$ under $f$ has distribution $\BE(a',Z',D')$. The
277: restriction property is straightforward; conformal invariance has
278: been first proved by \cite{levy40}; the later development of
279: stochastic calculus makes it a simple exercise.
280:
281: It is important for the previous paragraph that by Theorem 9.8
282: in the book of \cite{pommerenke}, conformal homeomorphisms from
283: the half plane $\HH$ or unit disk to regular domains extend
284: continuously to the boundary.
285:
286: \begin{remark}\lb{curve} \rm
287: One way to get such domains is the following. Let $E$ denote the
288: image of a curve $\gamma:[0,1]\rightarrow \CC$, and let $D$ be a
289: connected component of $\CC\setminus E$. Since local connectivity
290: is preserved under continuous maps, $E$ is locally connected. The
291: proof of Theorem 9.8 in \cite{pommerenke}, part (ii) $\rightarrow$
292: (iii) carries through without changes to show that in fact the
293: boundary of $D$ is also locally connected.
294: \end{remark}
295:
296: The strong Markov property says that the future of $B$ after a
297: stopping time $\tau$ given $\FF_\tau$ has distribution
298: $\BE(B(\tau),Z,D)$.
299:
300: The following theorem is due to \cite{cm83} (see
301: \cite{MR86d:60088} for a simpler proof).
302:
303: \begin{theorem}\label{cm}
304: There exists a universal constant $c$ so that the lifetime $\tau$
305: of a path distributed $\BE(a,Z,D)$ has $\ev\tau < c\;$area$(D)$.
306: \end{theorem}
307:
308: We say that a sequence $a_n\in D$ {\bf converges to $a\in \partial
309: D$ along a path} if there is a continuous curve $\gamma:(0,1)\to
310: D$ so that $a_n=\gamma(t_n)$ for some sequence $t_n$, and $\gamma$
311: is contained in the equivalence class that defines the prime end
312: $a$.
313:
314: In order to define a distance, we may extend paths defined on a
315: finite interval $[0,t]$ to be constant on $(-\infty,0]$ and
316: $[t,\infty)$. Let
317: \begin{eqnarray*}
318: \dist_0(\pi_1,\pi_2)&=&\sup_{t\in \RR}\|\pi_1(t)-\pi_2(t)\|,\\
319: \dist(\pi_1,\pi_2)&=&\inf_{\eps \in
320: \RR}\left(|\eps|+\dist_0(\theta_\eps\pi_1,\pi_2)\right).
321: \end{eqnarray*}
322: where $\theta_\eps$ denotes time shift. Let ``$\Rightarrow$''
323: denote convergence in distribution (in the case of paths with
324: respect to the metric ``$\dist$'').
325:
326: \begin{proposition}\lb{kezdet}
327: If $a_n\to a\in \partial D$ along a path, then $\BE(a_n,z,D)
328: \Rightarrow \BE(a,z,D).$
329: \end{proposition}
330:
331: \newcommand{\cdist}{\mbox{cdist}}
332: \begin{proof}
333: Consider the conformal homeomorphism $g:\HH \rightarrow D$ which
334: maps $0$ to the prime end $a$ (as determined by the path of
335: convergence of $a_n$). Since the domain $D$ is regular, we have $
336: g^{-1}(a_n)\rightarrow 0 $. Let
337: $$N_\eps=\{g(z)\ : \ z\in \HH, |z|<\eps\}.$$
338: Since the map $g$ is uniformly continuous in a neighborhood of
339: $0$, it follows that diam$(N_\eps)\rightarrow 0$ as $\eps \to 0$.
340:
341: If $a_n\in N_\eps$, then let $p_{n,\eps}$ be the probability that
342: independent paths with distributions $\BE(a_n,Z,D)$ and
343: $\BE(a,Z,D)$ do not intersect before their respective hitting
344: times $\tau_{n,\eps}$, $\tau_\eps$ of $N_\eps$. By conformal
345: invariance, this probability can be computed in the half-plane
346: image. Then it is well known that there exists $c,\gamma>0$ so
347: that
348: $$
349: p_{n,\eps}:=\| \mu_{n,\eps} - \mu_{\eps}\|_{TV} <
350: c(|g^{-1}(a)-g^{-1}(a_n)|/\eps)^\gamma.
351: $$
352: Indeed, one can consider times of hitting concentric circles of
353: radii $2^{-k}$. Between such times the probability that two
354: Brownian excursions intersect is bounded away from $0$. Thus
355: $p_{n,\eps}\rightarrow 0$ for $\eps$ fixed and $n\rightarrow
356: \infty$.
357:
358: Let $B$ have distribution $\BE(a,z,D)$, and define $B_n$ as having
359: distribution $\BE(a_n,z,D)$ and coupled to $B$ at its first
360: hitting time of the path of $B$. The event $C_n$ that this can be
361: done before either process has left $N_\eps$ has probability
362: $1-p_{n,\eps}$.
363:
364: Let $A_n$ be the event that $|\tau_{\infty,\eps}-\tau_{n,
365: \eps}|<\delta$. By Theorem \ref{cm} and Markov's inequality,
366: $\pr(A_n^c)<c_2\; $area$(N_{\eps})/\delta$, for some universal
367: constant $c_2$. Now if $A_n$ and $C_n$ hold, and $\eps<\delta$,
368: then $\dist(B,B_n)<3\delta$. Therefore
369: $$
370: \pr(\dist(B,B_n)> 3 \delta) \le c_2\;\mbox{area}(N_{\eps})/\delta
371: + p_{n,\eps} + \one(\delta\le \eps)
372: $$
373: and the right hand side can be made arbitrarily small by first
374: picking $\eps$ then letting $n\rightarrow \infty$. It follows
375: that $B_n$ converges to $B$ in probability and so in distribution.
376: \end{proof}
377:
378: \begin{corollary}\label{01H}
379: Let $Z_n\rightarrow {1}$ be sets of reals, and let $a_n\in \HH$,
380: $a_n \rightarrow 0$. Then $$\BE(a_n,Z_n,\HH) \Rightarrow
381: \BE(0,1,\HH).$$
382: \end{corollary}
383:
384: \begin{proof}
385: It suffices to prove this for the case when each $Z_n=\{z_n\}$ is
386: a single point, since the general case is a mixture of such
387: processes. Consider the rescaled versions $\BE(a_n/z_n,1,\HH)$ of
388: $\BE(a_n,z_n,\HH)$. These converge to $\BE(0,1,\HH)$ by the lemma,
389: and since the scaling factor converges to 1, so do the original
390: processes.
391: \end{proof}
392:
393: Let $\BE(a,z,D,K)$ denote the distribution of $\BE(a,z,D)$
394: conditioned to hit $K$. Say a sequence of sets $A_n\subset D$ {\bf
395: converges to $a$ along a path} if every sequence of points $a_n\in
396: A_n$ does.
397:
398: \begin{lemma}\label{slimhit}
399: Let $D$ be a regular domain, and consider regular subdomains $D_n
400: = D \setminus K_n$, where the $K_n \subset D$ are relatively
401: closed. Let $L_n \subset D_n$ closed, and let $a_n\in D_n$.
402: Suppose that
403: $$M_n:=K_n\cup L_n \cup \{a_n\} \rightarrow \{a\}\subset
404: \partial D \quad \mbox{along a path}.$$ Then
405: $$\BE(a_n,z,D_n,L_n) \cd
406: \BE(a,z,D).$$
407: \end{lemma}
408:
409: \begin{proof}
410: Let $W_n$ have distribution $\BE(a_n,z,D_n,L_n)$. The slightly
411: difficult part is to construct relatively closed sets $S_n
412: \rightarrow \{a\}$ separating $z$ from $M_n$ in $D$ so that
413: %
414: \bel{e1}
415: \pr(W_n \mbox{\ hits\ } S_n \mbox{ before }L_n)\rightarrow 0,
416: \ee
417: \vskip-1em
418: \bel{e2}
419: \sup_{s\in S_n} \pr(\BE(s,z,D) \mbox{ hits } K_n) \rightarrow 0.
420: \ee
421:
422: To complete the proof from here, let $\tau_n$ be the hitting time
423: of $S_n$ for $W_n$. Since $S_n \rightarrow \{a\}$, we have $\tau_n
424: \rightarrow 0$. If we condition $W_n$ to have done its duty of
425: hitting $L_n$ by time $\tau_n$, then $\{W_n(\tau_n+t),\ t\ge 0\}$
426: has the same distribution as $\BE(W(\tau_n),z,D_n)$. This, in turn
427: has the same distribution as $\BE(W(\tau_n),z,D)$ conditioned not
428: to hit $K_n$. In both cases, we are conditioning on events whose
429: probabilities converge to 1. Also, $\BE(W(\tau_n),z,D)$ is a
430: mixture of processes starting from points of $S_n$, so by
431: Proposition \ref{kezdet} it converges to $\BE(a,z,D)$. It follows
432: that the distribution of $W_n$ converges to $\BE(a,z,D)$, as
433: required.
434:
435: Now we proceed to find the sets $S_n$. By conformal invariance
436: of the probabilities involved, it suffices to do this in the
437: upper half plane with $a=0,\ z=\infty$. Here we use that our
438: domain has a curve boundary, so if in $\HH$ we have $S_n
439: \rightarrow 0$, then for its image under the conformal
440: homeomorphism $(0,\infty,\HH) \to (a,z,D)$ we have $S'_n
441: \rightarrow a$.
442:
443: Let $c_n\rightarrow \infty$ slow enough that we still have $c_n
444: M_n \rightarrow {0}$; if we construct $S_n$ that are bounded for
445: this rescaled problem, then scaling back will make $S_n
446: \rightarrow {0}$. So it suffices to construct bounded $S_n$. We
447: consider the uniformizing map $g_n$ which takes $D_n$ to the half
448: plane $\HH$. Assume that $g_n$ has hydrodynamic normalization, and
449: extend it to the boundary of $D_n$. The half-plane version of the
450: Caratheodory Kernel Theorem (see Theorem 1.8 in \cite{pommerenke})
451: implies that $g_n(z)\rightarrow z$ uniformly where it is defined.
452: So if we $L_n'=g_n(L_n)$, $a_n'=g_n(a_n)$ then
453: $L'_n\cup\{a'_n\}\rightarrow \{0\}$.
454:
455:
456: Let $S$ be the unit semicircle in $\HH$, and use the shorthand
457: $B_n=\BE(a'_n,\infty,\HH)$. We will show that
458: \bel{e5}
459: \pr(B_n \mbox{ hits } S \mbox{ before } L'_n\ |\ B_n \mbox{ hits }L'_n
460: )\rightarrow 0,
461: \ee
462: \vskip-1em
463: \bel{e6}
464: \sup_{s\in S}\pr (\BE(s,\infty,\HH) \mbox{\ hits\ } L'_n) \rightarrow 0.
465: \ee
466: Once we have this, we take $S_n= g_n^{-1}(S)$ so by uniform
467: convergence, $S_n \rightarrow S$. Thus $\{S_n\}$ is bounded, and
468: $(\ref{e1}, \ref{e2})$ follow from conformal invariance.
469:
470: The second statement (\ref{e6}) is obvious, and the first follows
471: since we can easily find semicircles $R_n \rightarrow \{0\}$ so
472: that
473: \bel{e7}
474: \sup_{w\in R_n}
475: \pr(\BE(w,\infty,\HH)\mbox{\ hits\ } L'_n) <
476: \pr(\BE(a'_n,\infty,\HH)\mbox{\ hits\ } L'_n).
477: \ee
478: The left hand side of (\ref{e5}) can be written as
479: \bel{e3}
480: \pr(\mbox{$B_n$ hits $S$ before $L'_n$ and hits $L'_n$})
481: \ /\ \pr(\mbox{$B_n$ hits }L'_n).
482: \ee
483: Using the Markov property and the topology of the setup the
484: numerator can be bounded above by
485: \begin{equation} \label{e4}
486: \sup_{w \in S} \pr(\BE(w,\infty,\HH)\mbox{ hits }R_n)
487: \sup_{w\in R_n}\pr(\BE(w,\infty,\HH)\mbox{ hits }L'_n)
488: \end{equation}
489: and so by (\ref{e7}), expression (\ref{e3}) is bounded above by
490: the first factor of (\ref{e4}), which converges to 0, proving
491: (\ref{e5}).
492: \end{proof}
493:
494:
495: \begin{lemma}\label{fathit}
496: Let $D$ be a regular domain, and let $K\subset D$ be a relatively
497: compact connected set containing at least two points. Let $z\in
498: \partial D$. Suppose that $K_n\subset D$ are relatively compact
499: and converge to $K$. Suppose that $a_n \rightarrow a$. Then
500: $$\BE(a,z,D,K_n)\cd \BE(a_n,z,D,K).$$
501: \end{lemma}
502:
503: \begin{proof}
504: Let $g$ be a bounded continuous test function for paths, and let
505: $B_n,\,B$ be distributed as $\BE(a_n,z,D)$ and $\BE(a,z,D)$
506: respectively.
507:
508: Note that for $\BE(a,z,D)$-almost all paths $\pi$ (1) the function
509: $f_n = \one(\pi\ hits\ K_n)$ converges to $f = \one(\pi\ hits\
510: K)$ and (2) $f$ is continuous at $\pi$. Thus by Proposition
511: \ref{kezdet} we have
512: \begin{eqnarray*}\ev(f_n(B_n)g(B_n)) &\rightarrow& \ev(f(B)g(B)),
513: \\
514: \ev f_n(B_n) &\rightarrow& \ev f(B).
515: \end{eqnarray*}
516: Note that the hypothesis of the lemma ensures that $\ev f(B)$ is
517: positive. Taking the ratio of the previous two limit statements we
518: get the desired result.
519: \end{proof}
520:
521: \subsection*{Hitting probabilities}
522:
523: Let $D$ be a domain, and let $a,z$ be points on $\partial D
524: \setminus A$ in the neighborhood of which the boundary is a
525: differentiable curve. Let $A$ be a hull in $\overline D$ not
526: containing $a,z$. We show the following equivalent version of
527: Proposition \ref{mainhit}.
528:
529: \begin{proposition}\lb{hit}
530: Let $f$ be a conformal homeomorphism that takes $D\setminus A$ to
531: $D$ and fixes $a,z$. Then
532: $$
533: \pr(\BE(a,z,D)\mbox{ avoids }A) = f'(a)f'(z).
534: $$
535: \end{proposition}
536:
537: \begin{proof}
538: Because of conformal invariance, it is sufficient to prove this in
539: the half-plane $\HH $ with $a=0$, $z=1$.
540:
541: The indicator function of the event above is continuous at almost
542: every path with respect to the excursion measure. Let $a_n=i/n$,
543: and $Z_n=(1,1+1/n)\subset \RR$. It suffices to show that as
544: $n\rightarrow \infty$ we have
545: \begin{equation}\label{ff1}
546: \pr(\BE(a_n,Z_n,\HH)\mbox{ avoids }A) \rightarrow f'(0)f'(1)
547: \end{equation}
548: since the measures on the left converge to the excursion measure
549: by Corollary \ref{01H}.
550: \begin{eqnarray}
551: \pr(\bm(a)\mbox{ hits }Z_n\mbox{ no later than }\partial
552: D)=h_{1/n}(i/n) = \nu n^{-2} + o (n^{-2}) \label{hit1}
553: \end{eqnarray}
554: Here $h_\eps$ is the harmonic function determined by the boundary
555: conditions $1$ on $(1,1+\eps)$ and $0$ elsewhere, and $\nu$ is
556: some fixed constant. The functions $h$ can be easily given
557: explicitly, but the formulas are unimportant, all we need is the
558: intuitive fact that as $\eps,z \rightarrow 0$ we have
559: $$h_\eps(z) = \nu\eps\Im z + o(\eps|z|)
560: $$
561: giving the second equality in (\ref{hit1}). Similarly, after a
562: conformal transformation by $f$
563: \begin{equation}\label{hit2}
564: \pr(\bm(a_n)\mbox{ hits }Z_n \mbox{ no later than }A\cup\partial
565: D)
566: \end{equation}
567: becomes
568: \begin{eqnarray}
569: \pr(\bm(f(a_n))\mbox{ hits }f(Z_n) \mbox{ no later than
570: }f(A\cup\partial \HH)) \quad \nonumber
571: \\=h_{f(1+\eps)-1}(f(i\eps)) = f'(0)f'(1)\nu
572: n^{-2} + o(n^{-2}) \nonumber
573: \end{eqnarray}
574: but the ratio of (\ref{hit2}) and (\ref{hit1}) gives (\ref{ff1}),
575: proving the claim.
576: \end{proof}
577:
578: \section{The cuttime filtration} \lb{s.ctf}
579:
580: Consider the excursion $\{B_t\}$ with distribution $\BE(a,z,D)$
581: where $D$ is a regular domain, $a,z\in \partial D$. Let $G$ denote
582: the set of {\bf cuttimes}, that is, the set of times for which
583: images of the future and past are disjoint; \cite{burdzy89} showed
584: that such times exist with probability 1. It is easy to check that
585: the set $G$ is measurable and it is a closed set a.s.
586:
587: It is clear that one cannot decide what the cuttimes are up to
588: time $t$ by only looking at the past of the process $\{B_t\}$. $G
589: \cap [0,t]$ is therefore not measurable with respect to the
590: standard filtration ${\mathcal F}_t$ of $\{B_t\}$. We therefore
591: introduce the {\bf cuttime filtration} $\mathcal G_t$ generated by
592: ${\mathcal F}_t$ and $G \cap [0,t]$.
593:
594: This enlargement of filtration may look large, but in fact it is
595: not in the following sense. Let $\Ghat_t$ denote the set of
596: cuttimes of the process $\{B_t\}$ restricted to the interval
597: $[0,t]$. If $\tau$ is the earliest time for which the set of
598: points $B[0,\tau]$ is revisited after time $t$, then we clearly
599: have
600: $$
601: G \cap [0,t] = \Ghat_t \cap [0,\tau].
602: $$
603: Since $\Ghat_t$ has Hausdorff dimension less than one
604: \citep{lawler96}, and the distribution of $B_\tau$ is absolutely
605: continuous with respect to harmonic measure on $B[0,t]$, we have
606: by Makarov's theorem that $\tau \notin \Ghat_t$ ${\mathcal
607: F}_t$-a.s. In particular, $G\cap[0,t]$ is determined by the
608: connected component $(g_t,g_t')$ of $[0,t] \setminus \Ghat_t$
609: containing $\tau$. The left endpoint of this interval, $g_t$, is
610: the last cuttime up to time $t$. Then given ${\mathcal F}_t$,
611: $g_t$ determines $\mathcal G_t$; its distribution is concentrated
612: on the countable set of points in $\Ghat_t$ that are isolated from
613: the right.
614:
615: Let $\rho B_s$ denote the connected component of $D\setminus
616: B[0,s]$ having $z$ as an interior or boundary point. By Remark
617: \ref{curve} if $D$ is a regular domain, then so is $\rho B_s$ for
618: every $s$. Let $\BE(a,z,D,K)$ denote the distribution of
619: $\BE(a,z,D)$ conditioned to hit $K$.
620:
621: \begin{proposition}[Markov property for cuttime filtration]
622: \lb{mp} Let $\tau$ be a ${\mathcal G}_t$-stopping time. Then the
623: distribution of $\{B_{\tau+t}, t\ge 0\}$ given $\mathcal G_\tau$
624: is $\BE(B_t,z,\rho B_{g_t}, B[g_t,g_t'])$.
625: \end{proposition}
626:
627: Note that, while for fixed time $\tau$, $B_{\tau}$ is an interior
628: point of $\rho B_{g_\tau}$ with probability one, there exists
629: stopping times for which $g_\tau=\tau$ a.s. An example is the
630: first cuttime after time 1. In these cases the ``hitting''
631: condition of the proposition holds trivially.
632:
633: \begin{proofof}{Proposition \ref{mp}} \
634:
635: Recall that $\GG_t$ is generated by $ \FF_t$ and $g_t$. Since the
636: distribution of $\{B_{t+s},s \ge 0\}$ given $\FF_t$ is
637: $\BE(B_t,z,D)$ by the Markov property, further conditioning on the
638: value of $g_t$ means that the future of the process $\BE(B_t,z,D)$
639: has to stay in $\rho B_{g_{\tau}}$, and hit
640: $B[g_{\tau},g'_{\tau}]$. This completes the proof for fixed time
641: $t$. Denote $L_t$ the set $B[g_{t},g'_{t}]$, and denote $\ev_t$
642: the distribution conditioned as above, that is $\BE(B_t,z,\rho
643: B_t, L_t)$.
644:
645: What we showed is equivalent to the following. If $X$ is a
646: bounded random variable on paths, and the time shift operator
647: $\theta_t$ is defined as $ (\theta_t B)(s)=B(s+t)$, then
648: $$
649: \ev (X\circ \theta_t \ |\ \GG_t) = \ev_t X.
650: $$
651: Thus, if $\tau$ is a $\GG_t$-stopping time concentrated on a
652: discrete set of values $\mathcal T$, then
653: \begin{eqnarray*}
654: \ev (X\circ\theta_\tau\ |\ \GG_\tau)
655: &=& \sum_{t\in \mathcal T} \ev (X\circ \theta_\tau\ \one(\tau=t) \ |\ \GG_\tau)
656: \ =\ \sum_{t\in \mathcal T} \ev (X\circ \theta_t\ \one(\tau=t) \ |\ \GG_t)
657: \\&=& \sum_{t\in \mathcal T} \one(\tau=t) \ev (X\circ \theta_t\ \
658: |\
659: \GG_t)
660: \ =\ \sum_{t\in \mathcal T} \one(\tau=t) \ev_t X
661: \ =\ \ev_\tau X,
662: \end{eqnarray*}
663: which completes the proof for this case. For an arbitrary
664: $\GG_t$-stopping time $\tau$ we take a discrete approximation by
665: letting $\tau_n$ be the first element of $ 2^{-n}\Z$ which is at
666: least $\tau$. Since the decision to stop at time $\tau_n$ is made
667: by time $\tau$, we have
668: $$
669: \ev(X\circ \theta_{\tau_n} \ |\ \GG_{\tau_n}) = \ev(X\circ
670: \theta_{\tau_n} \ |\ \GG_\tau).
671: $$
672: Since $\theta_t$ is a.s. continuous as $t \downarrow 0$, and $X$
673: is continuous, we have $X \circ \theta_{\tau_n} \rightarrow
674: X\circ\theta_{\tau}$ almost surely, and by the bounded convergence
675: theorem
676: $$
677: \ev(X\circ\theta_{\tau_n} \ | \ \GG_\tau) \rightarrow \ev(X\circ
678: \theta_\tau \ |\ \GG_\tau).
679: $$
680: Therefore it suffices to prove that $\FF_\infty$-a.s.
681: $\ev_{\tau_n} X \rightarrow \ev_\tau X$, or
682: \bel{cd}
683: \BE(B_{\tau_n}, z, \rho B_{g_{\tau_n}},L_{\tau_n}) \cd
684: \BE(B_{\tau}, z, \rho B_{g_\tau},L_{\tau}).
685: \ee
686: First we fix the path $B$ and examine what happens to $g_{\tau_n}$
687: and $g'_{\tau_n}$ in the limit. The definition $g_\tau= \sup (G
688: \cap [0,\tau])$ implies that $g_\tau$ is increasing in $\tau$ and
689: continuous from the right. Assume that $\tau\not=g_\tau$. Then
690: $\tau$ lies in a connected component of $G$ whose left endpoint is
691: $g_\tau$, so for all large $n>N_1$ we must have $g_{\tau_n}=
692: g_\tau$.
693:
694: Now assume also that $g'_\tau \not=\tau$. Then $B[g_\tau,
695: g_\tau']$ is disjoint from $B(g_\tau', \tau]$. Let $t$ be the
696: first time after $\tau$ that $B_t$ revisits the set
697: $B[g_\tau,g_\tau']$. Then for $\tau_n<t$ and $n>N_1$ we have
698: $g'_{\tau_n}=g'_\tau$. On the other hand, if $g'_\tau=\tau$, then
699: $ g'_{\tau_n}\le {\tau_n}$ implies that $g'_{\tau_n}\downarrow
700: g'_{\tau}$.
701:
702: So whenever $\tau\not=g_\tau$, we have $\rho B_{g_{\tau_n}} = \rho
703: B_{g_\tau}$ eventually, $L_{\tau_n} \rightarrow L_{\tau}$, and the
704: latter contains at least two points. By Lemma \ref{fathit} we get
705: $$
706: \BE(B_{\tau_n}, z, \rho B_{g_\tau},L_{\tau_n}) \cd
707: \BE(B_{\tau}, z, \rho B_{g_\tau},L_{\tau}),
708: $$
709: which implies (\ref{cd}).
710:
711: If $g_\tau=\tau$, then set $K_n:=B[g_{\tau},g_{\tau_n}]$. Note
712: that $K_n \cup L_{\tau_n} \cup \{B_{\tau_n}\} \subset B[\tau,
713: \tau_n]$ converges to $\{B_\tau\}$ along a path (i.e. a sub-path
714: of $B_t$). Therefore Lemma \ref{slimhit} yields (\ref{cd}).
715: \end{proofof}
716:
717: \section{Conformal maps, paths, semigroups} \lb{s.conformal}
718:
719: We begin this section with some notation and standard facts about
720: complex geometry. Most can be found in \cite{pommerenke},
721: \cite{lawlersle}, and best of all the upcoming book
722: \cite{lawlerbook}, which reviews complex geometry with an eye
723: towards applications in stochastic processes.
724:
725: If $A\subset \HH$, let $h_y(A)$ be harmonic measure, or the
726: hitting distribution of Brownian motion started at $y$ and stopped
727: at $A\cup \RR$. Let
728: $$h_\infty(A)=\lim_{y\rightarrow\infty} y\cdot
729: h_y(A)$$ Define
730: \begin{eqnarray}\label{capz}
731: \capz(A)&=&h_\infty(A),\\
732: \label{capo}
733: \capo(A)&=&\int_{\overline A} \Im(z) h_\infty(dz).
734: \end{eqnarray}
735: If $D$ is a subset of $\HH$ so that $\HH\setminus D$ is bounded,
736: then the capacity $\capz$ in $D$ can be defined analogously and is
737: denoted $\capz(D,A)$. Conformal invariance of harmonic measure
738: implies that if $f$ is a conformal homeomorphism $D\rightarrow
739: \HH$ satisfying $f'(\infty)=1$, then $\capz(D,A)=\capz(f(A))$. The
740: quantity $\capo$ will be referred to as half-plane capacity.
741:
742: If $A$ is bounded, $A \cup \RR$ is connected, and $f:\HH\setminus
743: A \to \HH$ is a conformal homeomorphism with hydrodynamic
744: normalization (i.e. $f(z)-z\to 0$ as $z\to\infty$), then we have
745: \bel{hyd}
746: f(z)=z+\capo(A)/z+O(1/z^2)
747: \ee
748: as $z\rightarrow \infty$. Such maps have the property that for all
749: $z\in \HH$
750: \bel{shiftdown}
751: \Im f(z)\le \Im(z).
752: \ee
753: Recall the following standard fact about complex geometry. There
754: exist a constant $c$ so that if $A\subset \overline \HH$ is
755: connected, then
756: %
757: \bel{capdia}
758: c^{-1} \diam(A)\le\capz(A)\le c\,\diam(A),
759: \ee
760: %
761: if $A\cup \RR$ is connected and $x+iy\in A$, then
762: %
763: \bel{capim}
764: y^2/4 \le \capo(A).
765: \ee
766: Also, if $A,A'$ are disjoint, then we have
767: %
768: \bel{suba}
769: \capo(A\cup A')\le \capo(A)+\capo(A')
770: \ee
771: %
772: This implies that if $A_1,A_2,A$ are disjoint, with $A_1\cup R$
773: and $\HH\setminus A_1$ connected, then
774: %
775: \bel{concap}
776: \capo(A_1 \cup A_2 \cup A)-\capo(A_1 \cup A) \le \capo(A_1\cup
777: A_2)-\capo(A_1)
778: \ee
779: %
780: We apply the conformal map $f:(\HH\setminus A_1,0,\infty)\to
781: (\HH\setminus A_1,0,\infty)$; then by additivity of $\capo$ the
782: desired inequality transforms to $\capo(f(A_2\cup A))-\capo(f(A))
783: \le \capo(f(A_2))$, which follows from \re{suba}.
784:
785: Let $\Phi$ denote the set of paths $\pi:[0,\tau]\rightarrow
786: \overline \HH$ which intersect the real line only at time $0$ and
787: $\pi(\tau)$ is in the boundary of $\rho\varphi$, the infinite
788: connected component of the complement of the image $\pi[0,\tau]$.
789:
790: To each $\pi \in \Phi$ we can associate the unique conformal
791: homeomorphism $f_\pi:\HH \rightarrow \rho_\pi$ with
792: hydrodynamic normalization. which fixes $\infty$, has
793: derivative $1$ there, and extends to the boundary $\RR$
794: continuously (see Remark \ref{curve}) and maps $0$ to $\pi(t)$.
795: Such a map has expansion
796: \bel{shiftexp}
797: f(z)=z + a_0 + az^{-1} +
798: O(z^{-2})
799: \ee
800: at $\infty$. The coefficient $a=\capo(\pi)$ behaves additively
801: under compositions of maps. It has the scaling property
802: \bel{scaling}
803: a(r \pi)=r^2 a(\pi).
804: \ee
805: Using the conformal maps, it is possible to compose two paths in
806: $\Phi$. Let $\pi\circ \pi'$ equal $\pi$ on $[0, \ell_\pi]$, and
807: equal the $f_\pi$-transform of $\pi'$ (see the definition below)
808: time-shifted by $\ell_\pi$ for the rest of the time interval.
809:
810: Thus $\Phi$ is a semigroup, and the map $a$ is a semigroup
811: homomorphism $\Phi \rightarrow [0,\infty)$.
812:
813: \begin{definition} \rm Let $\pi$ be a path in a regular domain $D_1$
814: (possibly starting and ending at $\partial D_1$), and let $f$
815: be a conformal homeomorphism $D_1 \rightarrow D_2$. If the
816: integral
817: %
818: \bel{timechange}
819: s(t)=\int_0^t |f'\circ \pi(\tau)|^2 d\tau
820: \ee
821: %
822: is finite for each $t$, then let $t(s)$ be the inverse of $s(t)$;
823: the function $f\circ \pi\circ t (s)$ is called the {\bf
824: $f$-transform} of $\pi(t)$.
825: \end{definition}
826:
827: \section{Independent increments and a local time for beads}
828: \lb{s.independent}
829:
830: Consider the process $B\sim\BE$, and let $G$ denote the set of
831: cuttimes. If $B(t)$ is on the boundary of $\rho B_t$, let $f_t$
832: denote the normalized conformal homeomorphism mapping $(\rho
833: B_t,B(t), \infty)$ to $(\HH,0,\infty)$. In this section, we
834: explore the following consequence of Proposition \ref{mp} and the
835: conformal invariance of Brownian excursion.
836:
837: \begin{corollary}[Independent increments at cuttimes]
838: \label{iip} Let $\tau$ be a ${\mathcal G}_t$-stopping time
839: supported on the set of cuttimes $G$. Then the $f_\tau$-map of the
840: future has distribution $\BE$.
841: \end{corollary}
842:
843: By a minor abuse of notation, let $a(t)$ denote the half-plane
844: capacity of $B[0,t]$.
845:
846: \begin{lemma} The set $a(G)$ is a closed regenerative set
847: with a scale-invariant distribution.
848: \end{lemma}
849: Recall that a random set $S$ has scale-invariant distribution for
850: every $c>0$ the sets $cS$ are identically distributed. It is
851: called regenerative if for every $\{\sigma[S \cap [0,t]];\; t\ge
852: 0\}$-stopping time $\nu$ which satisfies $\nu\in S$ a.s. the
853: translated set $(S-\tau) \cap [0,\infty)$ has the same
854: distribution as $S$.
855:
856: \begin{proof}
857: Closedness follows from the continuity of half-plane capacity and
858: since the set of cuttimes is closed. Scale-invariance follows from
859: the scale-invariance of Brownian excursion and the scaling
860: property $\re{scaling}$ of the half-plane capacity $a$.
861:
862: To check the regenerative property, assume that the stopping time
863: $\nu$ is as above. Then $\tau=a^{-1}(\nu)$ is a stopping time for
864: the cutpoint filtration $\GG$, and $\tau=a^{-1}(\nu)$ is supported
865: on cutpoints. Therefore by Corollary \ref{iip} the distribution of
866: the $f_\tau$-mapping of the future of $B$ after $\tau$ has
867: distribution $\BE$.
868:
869: In particular, since $a$ behaves additively under conformal
870: homeomorphisms, the distribution of $(a(G)-\nu)\cap[0,\infty)$
871: given the past of the excursion is the same as the distribution
872: of $a(G)$. Thus $a(G)$ is a regenerative set, as required.
873: \end{proof}
874:
875: By a result of \cite{kingman73}, the above implies that $a(G)$ is
876: the image of a stable subordinator. More precisely, there exists a
877: nondecreasing random function $(\lambda(a),\, a \ge 0)$ adapted to
878: the filtration $(\sigma(a(G) \cap [0,a]),\, a \ge 0)$ so that
879: $\lambda$ increases exactly on the set $a(G)$ its right-continuous
880: inverse $(a_\lambda,\, \lambda\ge 0)$ is a stable subordinator
881: with index $\alpha\in (0,1)$. The normalization can be chosen so
882: that the L\'evy measure (the intensity of the Poisson point
883: process of jumps) assigns mass $y^{-\alpha}$ to the interval
884: $(y,\infty)$ for all $y\ge 0$. We call $\lambda$ the {\bf bead
885: (local) time}.
886:
887: Let $B_{\rightarrow \lambda}$ denote the process $(B(t), 0\le t
888: \le t(\lambda))$ as an element of the semigroup $\Phi$. Recall
889: that a L\'evy process $X$ on a topological semigroup $\Phi$ is a
890: right continuous process with left limits with the property that
891: at any fixed time $t$, given the entire past $(X_s, 0\le s \le
892: t)$, the distribution of $X$ is the same as the past composed with
893: an independent copy of $X$. Since the inverse bead local times are
894: $\GG_t$-stopping times supported on $G$, Corollary \ref{iip}
895: immediately gives
896:
897: \begin{proposition}$(B_{\rightarrow \lambda},\, \lambda\ge 0)$ is a
898: $\Phi$-valued L\'evy process.
899: \end{proposition}
900:
901: Let $t(\lambda)$ be the time corresponding to local time
902: $\lambda$, i.e. the solution of $a(t(\lambda))=a_\lambda$.
903: Whenever the process $a_\lambda$ has a jump, the segment
904: $(B(t),\,t\in[t(\lambda-),t(\lambda+)])$ has no cutpoints, while
905: $t(\lambda-),\,t(\lambda+)$ are cuttimes. Let $\beta(\lambda)$
906: denote the $f$-mapping of this segment from $\rho B(t(\lambda-))$
907: to $\HH$. Then $\beta(\lambda)$ is an element of the semigroup
908: $\Phi$. When $a(\lambda)$ has no jump, we set $\beta(\lambda)$
909: equal some null state.
910:
911: Note the following deterministic fact. For an interval
912: $I=[\lambda_1,\lambda_2]$ the set $\beta(I)$ equals the set
913: $\beta'(I-\lambda_1)$ for the process which is the
914: $f_{a(\lambda_1)}$-image of the original one. Since the inverse
915: bead local times are stopping times, the independent increment
916: property of Corollary \ref{iip} implies that for non-overlapping
917: intervals $I$, the sets $\beta(I)$ are independent. We have shown
918:
919: \begin{proposition} $\beta$ is a
920: $\Phi$-valued Poisson point process.
921: \end{proposition}
922:
923: The {\bf Brownian bead measure} ($\BB$) is the $\sigma$-finite
924: intensity measure of the Poisson point process $\beta$. In simple
925: terms, for a set of paths $A$ the measure $\BB(A)$ equals the
926: expected number of elements of $A$ among the beads $\beta[0,1]$.
927:
928: \section{Properties of beads}\lb{s.properties}
929:
930: The goal of this section is to establish some simple properties of
931: the {\bf bead process}, i.e. the measure $\BB$.
932:
933: {\bf Scaling.} It follows from the scaling properties of $\BE$ and
934: half-plane capacity that for $c>0$ we have
935: $\BB(rA)=r^{-2a}\BB(A)$, where scaling a path by $c$ means scaling
936: in space by a factor $c$ and time accordingly. This implies that
937: the law $\BB$ can be decomposed as a product of measures on
938: ``shape space'' and ``size space''. Let $\beta$ be chosen
939: according to $\BB$ conditioned to have size $a$ at least $1$, and
940: rescale $\beta$ to have size 1. The resulting probability measure
941: $\nbb$, determines the shape of $\BB$, while the size is given
942: independently by the measure $d(y^{-\alpha})$.
943:
944: We now check that the Markov property for cuttimes and conformal
945: invariance imply a Markov property of beads.
946:
947: Note that the time of beads is only defined up to translation
948: (since the integral in the $f$-transform may not be finite). We
949: may pick a rule to set $t=t_1$ for some fixed number $t_1$ when
950: (and if) the bead grows to size 1 (which here is an arbitrary
951: positive number). Let $t_0$ be the starting time of the bead
952: (possibly $-\infty$), and let $A_t$ denote the image of the time
953: interval between $t_0$ and the first $t$-local cuttime after $t_0$
954: under $B$. Let $T\ge t_1$ be a stopping time with respect to the
955: canonical filtration generated by the past of the process.
956:
957: \begin{lemma}[Markov property of beads]\lb{mpb}
958: Under the measure $\BB$, the process $(B(T+s),\, s\ge 0)$ has the
959: same distribution as a process $B'$ distributed as
960: $\BE(B(T),\infty, \HH,A_T)$ and stopped at a random time $\tau$.
961: Here $\tau$ is the first time that
962: $$(B[t_0,\tau]\cup B'[0,\tau))\cap B'(\tau,\infty] =0.$$
963: \end{lemma}
964: %
965: \begin{proof}
966: It suffices to show this for the case $T=t_1$ (recall that the
967: bead starts at a time $t_0<t_1$). The general case follows from
968: first applying the lemma at $T=t_1$ and then using the Markov
969: property of Brownian excursion. More precisely, if $B'$ is as
970: given in the claim, and $T\ge t_1$ is a stopping time, then the
971: distribution of the future of $B$ after $T$ is just the
972: distribution of the future of $B'$ after $T$, and $B'\sim
973: \BE(B(T),\infty, \HH,A_T)$ stopped at $\tau$.
974:
975: For the $T=t_1$ case note that $\BB$ conditioned to have size at
976: least $1$ is by definition the following. $\BE$ is run until the
977: first time $S$ that $a(S)-a(g'(S))=1$ where $g'(S), g(S)$ are the
978: global cuttimes immediately before and after time $S$. Then $\BB$
979: is the $f_{g'(S)}$-map of $B(g'(S)+s,\,0\le s \le g(S)-g'(S))$.
980:
981: Therefore the statement of the lemma follows from the Markov
982: property of Proposition \ref{mp}, conformal invariance and the
983: fact that $f$-mappings preserve half-plane capacity.
984: \end{proof}
985:
986: Note that an example of such a stopping time is the first time
987: when $\BB$ hits the line with imaginary part $y$. This will be
988: used in Lemma \ref{lifetime} to show that in fact beads have
989: finite lifetime, and can be started at $t_0=0$. Once this has been
990: done, it is straightforward to extend Lemma \ref{mpb} to arbitrary
991: stopping times $T>t_0$.
992:
993: \begin{remark}[Beads determine the excursion]\lb{determine}\rm Another property shared by
994: beads and It\^o excursions is that the process $(\beta_\lambda,\,
995: \lambda \ge 0)$ determines the process $(B(t),\, t\ge 0)$. When
996: $B(t)$ is on the boundary of $B[0,t]$ (these are the so-called
997: pioneer points), the parameter $a_0$ in the conformal shift
998: \re{shiftexp} defines the ``horizontal'' location of $B(t)$. Given
999: $\beta$, it is straightforward to determine $a_0$ as a function of
1000: the half-plane capacity $a$ for each $a\ge 0$. This gives a
1001: L\"owner chain for the pioneer points of $B$ (see, for example,
1002: \cite{lawlersle} for definitions), and hence determines its outer
1003: boundary. Now assume that the outer boundaries of $B$ and $B'$
1004: agree, but they still differ in some way within a particular bead.
1005: Since $f$-mappings are one-to-one, then that bead has to be mapped
1006: by the same conformal shift to different points in the processes
1007: $\beta, \beta'$. This answers a question posed by an anonymous
1008: referee.
1009: \end{remark}
1010:
1011: The proof of the following simple fact is left to the reader.
1012:
1013: \begin{lemma} \lb{time}
1014: Let $A$ be a nonempty subset of $\RR\times (0,1]\subset \HH$ so
1015: that $A\cup \RR$ is connected. Let $B$ have distribution
1016: $\BE(z,\infty, \HH)$ conditioned to hit $A$ and stopped when this
1017: happens. Let $T_1$ be the time $B$ spends in the strip with
1018: imaginary parts between $1$ and $2$. There exists absolute
1019: constants $c,\gamma$ so that for all $t>0$ we have
1020: $$
1021: \pr(T>t) < c e^{-\gamma t}.
1022: $$
1023: \end{lemma}
1024:
1025: \begin{lemma}[Finite lifetime]\lb{lifetime}
1026: Let $T$ denote the lifetime of the process with distribution
1027: $\BB$. Then a.e. $T<\infty$.
1028: \end{lemma}
1029:
1030: \begin{proof}
1031: Let $B$ be a bead conditioned to hit $\Im z=y$. By the Markov
1032: property, the future of $B$ after this time is just $\BE(z,\infty,
1033: \HH)$ conditioned to hit a certain subset of the past. Let
1034: $T(y,2y)$ denote the time $B$ spends with imaginary part in this
1035: interval. By Lemma $\ref{time}$ and scale invariance
1036: $$
1037: \pr(T(y,2y)>ty^2) < ce^{-\gamma t}
1038: $$
1039: and therefore, for the unconditioned measure
1040: $$
1041: \BB(T(y,2y)>ty^2) < y^{-\alpha}c'e^{-\gamma t}.
1042: $$
1043: setting $y_n=2^{-n}$, and $t_n=c_1 n$ the right hand side becomes
1044: $ 2^{\alpha n}2^{-c_1\gamma n} $ which is summable for an
1045: appropriate choice of $c_1$. By the first Borel-Cantelli lemma
1046: (which also holds for $\sigma$-finite measures) we get that $\BB$
1047: a.e. for all large $n$
1048: $$
1049: T(2^{-n},2^{1-n}) \le c n 2^{-2n}
1050: $$
1051: summing this we get that for some ``random'' constant $K$ and all
1052: $y\le 1$ \bel{timeineq} T(0,y)< K y^2|\mbox{log}\,y| \ee But the
1053: Markov property and the existence of cutpoints for large times
1054: implies that after hitting $\Im(z)=1$, $B$ will have finite
1055: lifetime $T'$ a.s. Therefore $T\le T(0,1)+T'$ is finite $\BB$-a.e.
1056: as required.
1057: \end{proof}
1058:
1059: \section{The exponent giving the bead index }\lb{s.exponent}
1060:
1061: The goal of this section is to identify the index $\alpha$ of the
1062: stable process driving Brownian beads as an exponent for a large
1063: deviation event. Let $A(t)$ denote the event that the half-plane
1064: excursion $B$ has no cuttime between times $1$ and $t$.
1065:
1066: \newtheorem*{exponent.thm}{Theorem \ref{exponent}}
1067: \begin{exponent.thm}
1068: For large $t$, we have $\pr A(t) = t^{-\alpha + o(1)}$.
1069: \end{exponent.thm}
1070:
1071: \cite{math.PR/0302115} recently computed essentially the same
1072: exponent using generalized SLE processes, with the result
1073: $\alpha=1/2$. We will compute $\alpha$ directly by a simpler
1074: argument in the next section.
1075:
1076: Let $(a(\lambda),\lambda \ge 0)$ be a stable subordinator with
1077: index $\alpha$. We will use the following simple fact. It follows
1078: directly from \cite{bertoin} page 76 Proposition 2.
1079:
1080: \begin{fact}\label{bert} Let
1081: \[
1082: X =\min\left(a([0,\infty)) \cap [1,\infty) \right)
1083: \]
1084: then for some positive $c$ as $x\rightarrow \infty$ we have $
1085: \pr(X>x) \sim c x^{-\alpha}. $
1086: \end{fact}
1087:
1088: Let $A'(a)$ denote the event that there is no cuttime so that the
1089: past has half-plane capacity between $1$ and $a$. Fact \ref{bert}
1090: implies
1091: %
1092: \begin{equation}\label{alpha}
1093: \pr A'(a) \sim c a^{-\alpha}\quad\mbox{ as } a \rightarrow \infty.
1094: \end{equation}
1095: %
1096: In order to conclude Theorem \ref{exponent}, we only need to show
1097: that half-plane capacity and time are not too far from each other;
1098: in fact it suffices to show the following.
1099:
1100: \begin{lemma}
1101: We have $\pr(t/s<\capo\,B[0,t]<ts) \ge 1-ce^{-\gamma s}$.
1102: \end{lemma}
1103:
1104: \begin{proof}
1105: For a set $A$ in the plane, let $M_x$ and $M_y$ denote the sup of
1106: the absolute value of the projection of $A$ to the $x$ and $y$
1107: axes, respectively. If $A\subset \overline \HH$ contains zero and
1108: is connected, then by (\ref{capim}) and considering the half-plane
1109: capacity of a rectangle we get
1110: $$
1111: M_y^2/4 \le \capo(A) \le c\max(M_y^2,M_x).
1112: $$
1113: Now let $A=B[0,1]$. Then it is easy to check the following simple
1114: property of the maxima of Brownian motion and the 3-dimensional
1115: Bessel processes.
1116: $$
1117: \pr( 1/s<M_y^2 \le \max(M_y^2,M_x)<s) \ge 1-ce^{-\gamma s},
1118: $$
1119: and the claim follows by scale-invariance.
1120: \end{proof}
1121:
1122: \section{The value of the bead index $\alpha$}
1123: \lb{s.index}
1124:
1125: Consider the process $B\sim \BE$; in this section we will index
1126: $B$ either by time or bead local time $\lambda$; in the latter
1127: case we will stick to the notation $\lambda$. Let $\mathcal L$
1128: denote Lebesgue measure, let $\mu$ denote the random measure on
1129: the half plane given by
1130: $$
1131: \mu(A)={\mathcal L}(\lambda\,:\,B(\lambda)\in A),
1132: $$
1133: and let $\overline \mu(A):=\ev \mu(A)$. By the scaling of
1134: half-plane capacity $a(\lambda)$ and the scale-invariance of
1135: Brownian motion we get that for $r\ge 0$
1136: $$
1137: \omu(rA) = r^{2\alpha} \omu(A).
1138: $$
1139:
1140:
1141:
1142: \begin{lemma} \lb{strip}
1143: The $\omu$-measure of $\,\Strip_1=\RR \times (0,1]$
1144: is finite.
1145: \end{lemma}
1146: \begin{proof}
1147: Consider the random measure
1148: $$
1149: \mu(\Lambda,A)= {\mathcal L}(\lambda\in \Lambda\,:\,B(\lambda)\in
1150: A)
1151: $$
1152: and let $\omu(\Lambda,A):=\ev\mu(\Lambda,A)$. Let $\lambda$ be the
1153: first bead time at least $1$ so that $B(\lambda)\in \Strip_1$, and
1154: let $f$ denote the corresponding conformal shift. Let
1155: $p=\pr(\lambda<\infty)$; it is easy to check that $p<1$. For $n\ge
1156: 0$ we have
1157: \begin{eqnarray*}
1158: \ev(\mu([1,n+1],\Strip_1) \gi \GG_\lambda) &\le&
1159: \one(\lambda<\infty) \omu([0,n+1-\lambda],f(\Strip_1))\\&\le&
1160: \one(\lambda<\infty)\omu([0,n],\Strip_1)
1161: \end{eqnarray*}
1162: because of the cuttime Markov property (Proposition \ref {mp}) and
1163: the fact \re{shiftdown} that $f(\Strip_1)\subset \Strip_1$.
1164: Taking expected values gives
1165: $$\omu([1,n+1],\Strip_1) \le p
1166: \omu([0,n],\Strip_1)=:p \omu_n.$$ This yields the recursion
1167: $\omu_{n+1} \le 1 +p \omu_n$, which gives the bound $\omu_\infty
1168: \le 1/(1-p)$, as required.
1169: \end{proof}
1170:
1171: \begin{lemma}\lb{absolute}
1172: $\omu$ is absolutely continuous with respect to Lebesgue measure
1173: on $\HH$ with density bounded below and above on compacts.
1174: \end{lemma}
1175:
1176: \begin{proof}
1177: Let $A\subset \HH\setminus\{0\}$ be a closed set so that $A\cup
1178: \RR$ and $\HH \setminus A$ are connected, and let $f$ be the
1179: conformal homeomorphism $(\HH\setminus A,0,\infty)\to
1180: (\HH,0,\infty)$. Call such $f$ a subdomain map.
1181:
1182: Fix a path $B$ which avoids $A$, and consider its image $f(B)$.
1183: The cuttimes $g$ of $B$ can be parameterized by $a(g)$ i.e.
1184: $\capo$ of the past, as well as $a'(g)$ that is $\capo$ of the
1185: past of the image $f(B)$. Fact \re{concap} implies that if
1186: $g_1<g_2$ are two cuttimes, then
1187: $$
1188: a(g_2)-a(g_1) \le a'(g_2)-a'(g_1),
1189: $$
1190: in particular, beads are smaller when measured by $a'$ then when
1191: measured by $a$. Let $E\subset \HH \setminus A$ a generic Borel
1192: subset, and let $\mu'(E)$ denote $\mu(E)$ measured for the
1193: $f$-mapping of the process $B$. Since a.s. $\mu(E)$ can be
1194: computed as the $\eps\to 0$ limit of the rescaled number of beads
1195: of size at least $\eps$ starting at a cutpoint in $E$, it follows
1196: that we have $\mu'(f(E))\le \mu(E)$. By the restriction property,
1197: $$
1198: \omu(f(E)) = \ev\left(\mu'(f(E))|B\mbox{ avoids }A\right) \le
1199: \ev(\mu(E)|B\mbox{ avoids }A),
1200: $$
1201: and therefore
1202: $$
1203: \omu(f(E)) \le \ev(\mu(E))/\pr(B\mbox{ avoids }A) =
1204: \omu(E)(f'(0)f'(\infty))^{-1}.
1205: $$
1206: Let $D_r(z)$ denote the open disk of radius $r$ about $z$. Let $K$
1207: be a compact subset of $\HH$. If $z\in K$ and $w$ is sufficiently
1208: close to $0$, then there exists a subdomain map $f$ so that
1209: $f(z)=w$. Consider the set of points $w$ for which $f$ exists for
1210: all $z\in K$; this set contains a rescaled version $sK$ of $K$. By
1211: compactness, we may choose subset maps for each $z\in K,sw\in sK$
1212: so that
1213: \begin{eqnarray*}
1214: c_0&\le& f'(0)f'(\infty), \\
1215: D_{sr}(sw)&\subseteq &f(D_{c_2r}(z))\qquad\mbox{ for all }r<c_1
1216: \end{eqnarray*}
1217: with constants $c_0,c_1,c_2$ depending on $K$ only. Then
1218: \begin{eqnarray*}
1219: s^{2\alpha}\omu(D_{r}(w)) &=& \omu(D_{sr}(sw)) \\&\le&
1220: \omu(f(D_{c_2r}(z))) \\&\le&
1221: (f'(0)f'(\infty))^{-1}\omu(D_{c_2r}(z))\\&\le&
1222: c_0^{-1}\omu(D_{c_2r}(z)).
1223: \end{eqnarray*}
1224: This uniform bound implies the claim by standard arguments about
1225: absolute continuity.
1226: \end{proof}
1227:
1228: The following are needed for the proof of
1229: \newtheorem*{indexhalf.thm}{Theorem \ref{indexhalf}}
1230: \begin{indexhalf.thm}
1231: \indexhalfthm
1232: \end{indexhalf.thm}
1233:
1234: Let $D_r=D_r(i)$ denote the open disk of radius $r$ about $i$. Let
1235: $\tau$ be the (possibly infinite) first hitting time of $D_r$ for
1236: the process $B\sim\BE$. Let $B^{(1)}$ denote the path image
1237: $B[0,\tau]$. Recall the definition of the capacity $\capz(D,A)$ of
1238: $A$ from $\infty$ in $D$ from Section \ref{s.conformal}, and use
1239: the shorthand $\capz(t,A)= \capz(\rho B_t,A)$. Let
1240: $K_r=\capz(\tau,D_r)$, and recall the notation $x_r \asymp y_r$
1241: for the existence of a constant $c>0$ so that $c^{-1}y_r \le
1242: x_r\le c y_r$ (here for all small $r$).
1243:
1244: \begin{proposition}\label{zbounds}
1245: As $r\to 0$ we have $ \omu(D_r) \asymp |\mbox{\rm log } r|\ev
1246: K_r^{1+2\alpha}. $
1247: \end{proposition}
1248:
1249: Let $B^{(2)}$ denote the path after the last exit from $D_r$. Let
1250: $Z'_r=\pr (B^{(1)} \nis B^{(2)}\,|\,\FF_\tau)$, where ``$\nis$''
1251: denotes ``does not intersect''. Let $Z_r$ denote the probability
1252: given $\FF_\tau$ that the image of an independent process
1253: $\BE(\infty, D_r,\HH)$ does not intersect $B^{(1)}$; recall that
1254: this image can be defined via conformal mapping of $\BE$ started
1255: at a more conventional boundary point.
1256:
1257: \begin{lemma}\lb{sizes} We have
1258: $|\mbox{\rm log}\ r| K_r \asymp Z_r \asymp Z'_r$, with
1259: deterministic constants.
1260: \end{lemma}
1261:
1262: \begin{proof}
1263: We will use the time-reversal property of Brownian excursion: if
1264: $B\sim \BE(a,z,D)$ and ends at random time $\tau$, then
1265: $(B(\tau-t),\, t\in[0,\tau])\sim \BE(z,a,D)$. This, as well as
1266: conformal invariance, can be used to define the image of
1267: $\BE(\infty,z,\HH)$ (together with a time-parameterization
1268: starting at $-\infty$, but we won't need this).
1269:
1270: The quantities in question are given by the measures of paths that
1271: do not intersect $B^{(1)}$ before hitting $D_r$ under the measures
1272: $|\mbox{log}\ r|\BE(\infty,\RR,\HH)$, $\BE(\infty, D_r,\HH)$, and
1273: $\BE(\infty,B(\tau),\HH)$, respectively. It follows from the
1274: definition of $\BE$ that the distribution of these paths up to the
1275: hitting time of $D_r$ given the hitting position agree. It is easy
1276: to check that the in each case hitting position has a smooth
1277: density bounded below and above with respect to uniform measure on
1278: $\partial D_r$. The $|\mbox{log}\ r|$ normalizing factor is
1279: necessary so that the measure of paths hitting $D_r$ is bounded
1280: below and above by constants as $r\to 0$.
1281: \end{proof}
1282:
1283: \begin{proofof}{Theorem \ref{indexhalf}}
1284: By the Lemma \ref{absolute}, as $r\rightarrow 0$ we have
1285: $$
1286: \omu(D_r) \asymp r^2.
1287: $$
1288: We have
1289: $$
1290: \ev K_r^\gamma \asymp |\mbox{log }r|^{-\gamma} \ev Z_r^\gamma =
1291: r^{\xi(1,\gamma)+o(1)},$$ where the first approximation follows
1292: from Lemma \ref{sizes}, and the second is one definition of the
1293: intersection exponent. More precisely, we should consider the
1294: analogue of $Z_r$ for $B^{(1)'}$, which has distribution
1295: $\BE(0,D_r,\HH)$, but this is absolutely continuous with density
1296: bounded above and below with respect to the distribution of
1297: $B^{(1)}$ given that it hits $D_r$ (see the proof of Lemma
1298: \ref{sizes}).
1299:
1300: Thus by Proposition \ref{zbounds} we get $
1301: r^{\xi(1,1+2\alpha)+o(1)} = r^2,$ and the theorem follows by the
1302: monotonicity of $\xi(1,\cdot)$ and the known value $\xi(1,2)=2$
1303: (see \cite{lawler95} for these properties of $\xi$).
1304: \end{proofof}
1305:
1306: \begin{proofof}{Proposition \ref{zbounds}}\ \\
1307: \indent{\bf Upper bound.} We write
1308: $$
1309: \ev(\mu(D_r)\,|\,\FF_\tau) = \ev(\mu(D_r) \,|\, B^{(1)}\nis
1310: B^{(2)}, \FF_\tau) \pr (B^{(1)} \nis B^{(2)}\,|\,\FF_\tau).
1311: $$
1312: Let $f$ denote the conformal shift at the first cuttime $g$ after
1313: time $\tau$. We have
1314: $$K_r=\capz(\tau,D_r)\ge \capz(g,D_r)=\capz(f(D_r)).
1315: $$
1316: Here, by slight abuse of notation, $f(D_r)$ denotes the image of
1317: the part of $D_r$ on which $f$ is defined. By \re{capdia} the
1318: above implies that $f(D_r)\subset \RR\times (0,cK_r].$ By scaling
1319: and Lemma \ref{strip} we have $\omu(\RR\times
1320: (0,K_r/2])=cK_r^{2\alpha}.$ Thus by the cuttime Markov property
1321: (Proposition \ref{mp}) we get
1322: $$
1323: \ev(\mu(D_r) \,|\, B^{(1)}\nis B^{(2)}, \FF_\tau) =
1324: \ev(\omu(f(D_r))\,|\,\FF_\tau) \le cK_r^{2\alpha}.
1325: $$
1326: Here and in the sequel $c$ denotes a constant whose value may
1327: change by line to line. By Lemma \ref{sizes}
1328: $$
1329: \ev(\mu(D_r) \gi \FF_\tau) \le c Z'_rK_r^{2\alpha} \le c|\mbox{log
1330: }r| K_r^{1+2\alpha},
1331: $$
1332: which implies the upper bound in Proposition \ref{zbounds}.
1333:
1334: \bigskip
1335: {\bf Lower bound.} Let $A_1$ denote the event that $B[0,\tau]$
1336: does not intersect the set $ \{w\in D_{2r}\,:\,\arg(w-i)\in
1337: (\pi/4,7\pi/4)\}.$ It is easy to check (see, for example
1338: \cite{lwssharp}) that for all $\gamma\in [1,3]$ and $r<1/2$ we
1339: have
1340: \bel{za1}
1341: \ev(Z_r^\gamma;\,A_1) \ge c\ev Z_r^\gamma.
1342: \ee
1343: Consider the event
1344: $$
1345: A=\{B(g)\in D_{r},\, B[0,g]\nis D_{r/2},\, \capz(g,D_{r/32})\ge
1346: cK_r \}.
1347: $$
1348: This implies $B^{(1)}\nis B^{(2)}$, and it is easy to check that
1349: there is an absolute constant $c_1$ so that
1350: $$
1351: \pr(A\,|\,B^{(1)}\nis B^{(2)},\FF_\tau) \ge c_1 \one_{A_1}.
1352: $$
1353: We have
1354: \begin{eqnarray*}
1355: \ev( \mu(D_{r}) \gi \FF_\tau) &\ge&
1356: \ev(\mu(D_{r});\, A \gi \FF_\tau)\one_{A_1} \\&\ge&
1357: \ev(\mu(D_{r}) \gi A,\FF_\tau)
1358: \pr(A \gi B^{(1)}\nis B^{(2)}, \FF_\tau)
1359: \pr(B^{(1)}\nis B^{(2)}\,|\, \FF_\tau)\one_{A_1}.
1360: \end{eqnarray*}
1361:
1362: By the cuttime Markov property (Proposition \ref{mp}) we get
1363: $$
1364: \ev(\mu(D_{r}) \gi A,\FF_\tau)
1365: = \ev(\omu(f(D_{r})) \gi A,\FF_\tau).
1366: $$
1367: The event $A$ implies that the domain of $f$ contains $D_{r/2}$,
1368: and therefore by the K\"obe quarter theorem $f(D_{r/2})$ contains
1369: a ball $D'$ of radius $|f'(i)|r/8$ centered at $f(i)$. By the same
1370: theorem applied to $f^{-1}$, we get that $f^{-1}(D')$ contains
1371: $D_{r/32}$. By monotonicity and conformal invariance of harmonic
1372: measure, we get
1373: $$
1374: \mbox{diam}(D') \ge c\capz(D')=c\capz(g,f^{-1}(D')) \ge
1375: c\capz(g,D_{r/32}) \ge c' K_r,
1376: $$
1377: where the last inequality is an assumption in $A$. Since
1378: $f(\partial D_r)$ contains an curve $K$ that separates $0$ and
1379: $D'$ from $\infty$ in $\HH$, we have
1380: $$
1381: \dist(D',0)\le \mbox{diam}(K)\le c\,\capz(K)\le
1382: c\,\capz(f(D_r))\le cK_r.
1383: $$
1384: When $K_r=1$, this implies that $\omu(D')$ is bounded below by a
1385: constant. The scaling property then implies that in general
1386: $\omu(D')\ge cK_r^{2\alpha}$. Putting all the above together and
1387: using Lemma \ref{sizes} we get
1388: $$ \ev (\mu(D_{r})\,|\,\FF_\tau) \ge c
1389: Z'_r K_r^{2\alpha}\one_{A_1} \ge c |\mbox{log }
1390: r|K_r^{1+2\alpha}\one_{A_1}
1391: $$ and the bound follows from (\ref{za1}).
1392: \end{proofof}
1393: \subsection*{Open questions and conjectures}
1394:
1395: \begin{question}\rm
1396: The parameter $a_0$ in the conformal shift \re{shiftexp} defines
1397: the ``horizontal'' location of a cutpoint. It follows from our
1398: proofs that the process $((a(\lambda),a_0(\lambda)),\, \lambda \ge
1399: 0)$ is a 2-dimensional L\'evy process stable under scaling with
1400: exponents $1/2$ and $1$ in the two respective coordinates. In
1401: particular $(a_0(\lambda),\, \lambda>0)$ is a Cauchy process. What
1402: is the joint distribution of the two processes? (This question may
1403: be better considered in the $\mbox{SLE}_6$ framework.)
1404: \end{question}
1405:
1406: \begin{question}\rm
1407: Is it possible to interpret bead local time as a local time at
1408: zero of some process? This may be better answered by considering a
1409: version of $\mbox{SLE}_6$.
1410: \end{question}
1411:
1412: \begin{conjecture}\rm
1413: Consider the $\sigma$-finite bead process conditioned to survive
1414: up to distance $r$ (or time $r^2$, or imaginary part $r$), and let
1415: $r\rightarrow \infty$. The limiting process exists and has the
1416: restriction property. The exponent (Werner, personal
1417: communication) should equal 2 (see \cite{lwsrest} for
1418: definitions).
1419: \end{conjecture}
1420:
1421: \begin{question}\rm
1422: The bead time $\lambda$ defines a random measure on the
1423: half-plane. The closed support of this measure is the set of
1424: cutpoints. Is it possible to derive the Hausdorff dimension of the
1425: set of cutpoints using this measure, or more generally, using
1426: beads? This would give a new, conceptual proof for the value of
1427: the intersection exponent $\xi(1,1)=5/4$.
1428: \end{question}
1429:
1430: \bigskip
1431: \noindent {\bf Acknowledgments.} The author thanks Jim Pitman for
1432: stimulating remarks and recommending the reference \cite{gp80},
1433: which can also be used to construct bead local time. He also
1434: thanks Oded Schramm and Wendelin Werner for inspiring discussions
1435: and conscientious referee for several important remarks and
1436: corrections.
1437: \bibliography{bbb}
1438:
1439: \end{document}
1440: