1: %Revised on September 8, 2005
2:
3: \magnification=\magstep1
4:
5: \baselineskip=1.3\baselineskip
6:
7: \font\tenmsb=msbm10 \font\sevenmsb=msbm7 \font\fivemsb=msbm5
8: \newfam\msbfam
9: \textfont\msbfam=\tenmsb \scriptfont\msbfam=\sevenmsb
10: \scriptscriptfont\msbfam=\fivemsb
11: \def\Bbb#1{{\fam\msbfam\relax#1}}
12:
13: \def\bone{{\bf 1}}
14: \def\vphi{\varphi}
15: \def\eps{\varepsilon}
16: \def\R{{\Bbb R}}
17: \def\Z{{\Bbb Z}}
18:
19: \def\P{{\bf P}}
20: \def\E{{\bf E}}
21:
22: \def\<{\langle}
23: \def\>{\rangle}
24:
25: \def\cG{{\cal G}}
26: \def\cE{{\cal E}}
27: \def\cV{{\cal V}}
28: \def\cS{{\cal S}}
29: \def\cB{{\cal B}}
30: \def\cT{{\cal T}}
31: \def\cU{{\cal U}}
32: \def\cW{{\cal W}}
33: \def\cA{{\cal A}}
34: \def\cL{{\cal L}}
35: \def\n{{\bf n}}
36: \def\e{{\bf e}}
37: \def\d{{\bf d}}
38: \def\v{{\bf v}}
39: \def\w{{\bf w}}
40: \def\z{{\bf z}}
41: \def\wt{\widetilde}
42: \def\wh{\widehat}
43: \def\prt{{\partial}}
44: \def\df{{\mathop {\ =\ }\limits^{\rm{df}}}}
45: \def\weak{{\mathop {\ \Rightarrow\ }}}
46: \def\nweak{{\mathop {\ \not\Rightarrow\ }}}
47: \def\ol{\overline}
48: \def\dist{{\mathop {\rm dist}}}
49: \def\vol{{\mathop {\rm vol}}}
50: \def\lam{\lambda}
51: \def\la{{\langle}}
52: \def\ra{{\rangle}}
53:
54: \def\qed{{\hfill $\square$ \bigskip}}
55: \def\sqr#1#2{{\vcenter{\vbox{\hrule height.#2pt
56: \hbox{\vrule width.#2pt height#1pt \kern#1pt
57: \vrule width.#2pt}
58: \hrule height.#2pt}}}}
59: \def\square{\mathchoice\sqr56\sqr56\sqr{2.1}3\sqr{1.5}3}
60:
61: \input epsf.tex
62: \newdimen\epsfxsize
63: \newdimen\epsfysize
64:
65:
66: \centerline{\bf SHY COUPLINGS}
67: \footnote{$\empty$} {\rm Research partially supported by
68: NSF grant DMS-0303310 (KB and ZC).}
69:
70: \vskip 0.3truein
71:
72: \centerline{\bf Itai Benjamini{\rm,} Krzysztof
73: Burdzy {\rm and} Zhen-Qing Chen}
74:
75:
76: \vskip0.4truein
77:
78: {\narrower
79:
80: \noindent
81: {\bf Abstract}.
82: A pair $(X,Y)$ of Markov processes is called a Markov
83: coupling if $X$ and $Y$ have the same transition probabilities
84: and $(X,Y)$ is a Markov process.
85: We say that a coupling is ``shy'' if
86: there exists a (random) $\eps>0$ such that
87: $\dist(X_t,Y_t) >\eps$ for all $t\geq 0$.
88: We investigate whether shy couplings
89: exist for several classes of Markov processes.
90:
91: }
92:
93: \vskip0.5truein
94:
95: \noindent {\bf 1. Introduction}. The proofs of the main theorems
96: in two recent papers, [BC2] and [BCJ], contained arguments showing
97: that certain processes come arbitrarily close to each other, at
98: least from time to time, as time goes to infinity, with
99: probability one. The proofs were based on ideas specific to the
100: models and were rather tedious. We decided to examine several
101: classes of Markov processes in order to determine the conditions
102: under which there exists a pair of Markov processes defined on the
103: same probability space such that each marginal process has the
104: same transition probabilities and the two processes do not come
105: close to each other at any time. Although we do not have a
106: complete solution to this problem, we offer a number of results
107: whose diversity points to a rich theory. Some of our theorems,
108: examples and techniques may have interest of their own.
109:
110: We will focus on two classes of processes---reflected Brownian
111: motions on Euclidean domains and Brownian motions on graphs. The
112: second class of processes is really discrete in nature, in the
113: sense that similar techniques work for random walks on graphs. We
114: chose these classes of processes because similar processes
115: appeared in our research in the past.
116:
117: For a general overview of coupling techniques, see [L].
118:
119: The rest of the paper is organized as follows. We present basic
120: definitions and elementary examples in Section 2. Section 3 is
121: devoted to Brownian motions on graphs. We show that there exists a
122: shy coupling for Brownian motions on
123: a graph if all its vertices have degree 3 or higher.
124: Four examples are also given to illustrate the case when the graph
125: has some vertices of degree one.
126: Section 4 deals with reflected Brownian motions on Euclidean
127: domains, showing that there exist no shy couplings on $C^1$-smooth
128: bounded strictly convex domains.
129:
130:
131: \bigskip
132: \noindent {\bf 2. Preliminaries and elementary examples}. Unless
133: specified otherwise, all pairs of processes $(X, Y)$ considered in
134: this paper will be ``Markov couplings,'' i.e., they will satisfy
135: the following assumptions.
136: \item{(i)} $\{X_t, t\geq 0\}$, $\{Y_t,
137: t\geq 0\}$ and $\{(X_t,Y_t), t\geq 0\}$ are Markov, and the
138: transition probabilities for $X$ and $Y$ are identical.
139: \item{(ii)} The distribution of $\{X_t, t\geq s\}$ conditional on
140: $\{(X_s, Y_s) = (x,y)\}$ is the same as the distribution of
141: $\{X_t, t\geq s\}$ conditional on $\{X_s = x\}$, for all $x,y $
142: and $s$.
143:
144: \medskip
145: Our definition of a Markov coupling is slightly different from
146: similar concepts in the literature. One could investigate the
147: question of whether our results hold for ``couplings'' defined in
148: other ways, for example, whether condition (ii) is essential.
149: However, we feel that there are more exciting open problems in
150: this area---see the end of Section 4.
151:
152: The following elementary discrete-time example shows that there
153: exist couplings that satisfy (i) but do not satisfy (ii).
154:
155: \bigskip
156: \noindent {\bf Example 2.1}. We take $\{0,1\}$ as the state space
157: of a discrete time Markov process and we let $\{X_k, k\geq 0\}$ be
158: a sequence of i.i.d. random variables with
159: $\P(X_k=0)=\P(X_k=1)=1/2$. We will define a process $\{Y_k, k\geq
160: 0\}$ with the same distribution as $\{X_k, k\geq 0\}$. We let
161: $Y_0$ be independent of $\{X_k, k\geq 0\}$. For $k\geq 1$, we
162: construct $Y_k$ so that $\P(Y_k=0 \mid X_{k-1} =0)=0.7$ and
163: $\P(Y_k=1 \mid X_{k-1} =1)=0.7$. Moreover, for every $k\geq 1$, we
164: make $Y_k$ independent of $X_j$'s for $j< k-1$. It is elementary
165: to check that $\{X_k, k\geq 0\}$, $\{Y_k, k\geq 0\}$ and
166: $\{(X_k,Y_k), k\geq 0\}$ are Markov but for $j\geq 1$, the
167: distribution of $\{Y_k, k\geq j\}$ conditional on $\{Y_{j-1} =0\}$
168: is not the same as the distribution of $\{Y_k, k\geq j\}$
169: conditional on $\{(X_{j-1},Y_{j-1}) =(0,0)\}$.
170:
171: \bigskip
172:
173: We will assume that the state space $\cS$ for Markov processes $X$
174: and $Y$ is metric and we will let $\d$ denote the metric. The open
175: ball with center $x$ and radius $r$ will be denoted $\cB(x,r)$.
176: The shortest path between two points in $\cS$ will be called a
177: geodesic. For some pairs of points, there may be more than one
178: geodesic joining them.
179:
180: \bigskip
181: \noindent {\bf Definition 2.2}. A coupling $(X,Y)$ will be called
182: shy if one can find two distinct points $x$ and $y$ in the state
183: space with
184: $$ \P \left( \inf_{0\leq t < \infty}
185: \d(X_t, Y_t)>0 \mid X_0=x, Y_0=y \right)>0.
186: $$
187:
188: \bigskip
189:
190: Note that the term ``shy coupling'' is a label for a family of
191: Markov transition probabilities.
192:
193: We proceed with completely elementary examples of shy and non-shy
194: couplings.
195:
196: \bigskip
197: \noindent {\bf Examples 2.3}. (i) Let $X$ be a Brownian motion in
198: $\R^d$ and let $0\ne y\in \R^d$ be a fixed vector. Let $Y_t = X_t
199: + y$ for all $t\geq 0$. Then $(X,Y)$ is a shy coupling.
200:
201: (ii) Let $X$ be a Brownian motion on the unit circle in $\R^2$ and
202: let $\theta \in (0,2\pi)$ be a fixed number. We define $Y_t$ using
203: complex notation, $Y_t = e^{i\theta} X_t$ for all $t\geq 0$. Then
204: $(X,Y)$ is a shy coupling.
205:
206: (iii) The last two examples can be easily generalized to a wide
207: class of Markov processes on spaces $\cS$ with a group structure.
208: If there is a group element $a\ne 0$ such that $a+X$ is a Markov
209: process having the same transition probabilities as $X$ and
210: $\inf_{b\in \cS} \d(b, a+b) >0$ then $(X,a+X)$ is a shy coupling.
211:
212: (iv) Let $X$ and $Y$ be independent Brownian motions in $\R^d$.
213: Then $(X,Y)$ is a shy coupling if and only if $d\geq 3$.
214:
215: \medskip
216: In the sequel, for $a, b \in \R$, $a\wedge b:=\min \{a, \, b\}$,
217: $a\vee b:=\max\{a, \, b\}$ and $a^+:=a\vee 0$. For $a>0$, $[a]$
218: denotes the largest integer that does not exceed $a$.
219:
220: \bigskip
221: \noindent {\bf 3. Brownian motion on graphs}. In this section, we
222: will consider processes whose state space is a finite or infinite
223: graph. More precisely, let $\cG = (\cV,\cE)$ be a graph, where
224: $\cV$ is the set of vertices and $\cE$ is the set of edges. We
225: will assume that all vertices have a finite degree, i.e., for
226: every vertex there are only a finite number of edges emanating
227: from this vertex, but we do not assume that this number is bounded
228: over the set of all vertices. We allow an edge to have both
229: endpoints attached to one vertex. Every vertex will be attached to
230: at least one edge.
231: We will identify edges with finite open line
232: segments (connected subsets of $\R$), with finite and strictly
233: positive length, and we will identify vertices with topological
234: endpoints of edges. In this way, we can identify the graph $\cG$
235: with a metric space $(\cS, \d)$, where $\cS = \cE \cup \cV$, and
236: $\d(x,y)$ is the shortest path between $x$ and $y$ along the edges
237: of the graph. We will assume that the length of any edge is
238: bounded below by $r_0>0$.
239:
240: Next we will construct ``Brownian motion'' $X$ on $\cS$. See [FW]
241: for a definition of a general diffusion on a graph. We leave it to
242: the reader to check that our somewhat informal description of the
243: process is consistent with the rigorous construction given in
244: [FW]. By assumption, our process will be strong Markov. Suppose
245: that $x\in e \in \cE$ and $x$ is not an endpoint of $e$. Recall
246: that $e$ can be identified with a line segment, say, $e=[0,y]$.
247: Then $x\in (0,y)$. If $X_0=x$, then the process $X$ evolves just
248: like the standard one-dimensional Brownian motion until the exit
249: time from $(0,y)$. Next suppose that $x\in \cS$ is a vertex. Then
250: there are $n$ edges $e_1, e_2, \dots, e_n$, attached to $x$, with
251: $n\geq 1$. Choose a small $r>0$ such that the ball $\cB(x,r)$
252: consists of line segments $I_j$, $j=1,2,\dots, k$, which are
253: disjoint except that they have one common endpoint $x$. Note that
254: $n \leq k \leq 2n$, but not necessarily $k=n$, because some edges
255: may have both endpoints at $x$. We will describe the evolution of
256: $X$ starting from $x$ until its exit time from $\cB(x,r)$.
257: Generate a reflected Brownian motion $R$ on $[0,\infty)$, starting
258: from 0, and kill it at the first exit time from $[0,r]$, denoted
259: $T_r$. Label its excursions from 0 with numbers $1, 2, \dots, k$,
260: in such a way that every excursion has a label chosen uniformly
261: from $\{1, \cdots, k\}$ and independently of all other labels.
262: Then we define $X_t$ for $t\in[0,T_r]$ so that $\d(X_t, x) = R_t$
263: and $X_t \in I_j$, where $j$ is the label of the excursion of $R$
264: from 0 that straddles $t$ (if $R_t=0$ then obviously $X_t =x$).
265: This defines the process $X_t$ until its exit time from
266: $\cB(x,r)$. What we said so far and the strong Markov property
267: uniquely define the distribution of $X$. Note that when the degree
268: of a vertex $x$ is 1 then $X$ is best described as a process
269: reflected at $x$. The process $X$ spends zero amount of time at
270: any vertex.
271:
272: Recall that we have assumed that the length of all edges is
273: bounded below by $r_0>0$. Under this assumption, the process
274: cannot visit an infinite number of vertices in a finite amount of
275: time. Hence, the above construction defines a process for all
276: $t\geq 0$. Another consequence of the assumptions that all edges
277: have length greater then $r_0$ and all vertices have finite degree
278: is that for any two points in $\cS$ there is only a finite number
279: of geodesics joining them. It is clear from our construction that
280: vertices of degree 2 will play no essential role in the paper and
281: can be ignored. So we will assume without loss of generality that
282: there are no vertices of degree 2.
283:
284: \bigskip
285: \noindent{\bf Theorem 3.1}. {\sl If all vertices of $\cG$ have
286: degree 3 or higher then there exists a shy coupling for Brownian
287: motions on $\cS$.}
288:
289: \bigskip
290: \noindent{\bf Proof}. We will construct a coupling $(X,Y)$ of
291: Brownian motions on $\cS$ such that $X$ and $Y$ move in an
292: independent way when they are far apart and they move in a
293: ``synchronous'' way when they are close together. Clearly,
294: independent processes do not form a shy coupling on a finite
295: graph. Remark 3.2 below explains why it is hard, perhaps
296: impossible, to construct a ``synchronous'' shy coupling.
297:
298: For any $x,y\in \cS$ with $\d(x,y) > r_0/4$, we will define
299: $(X_t,Y_t)$ starting from $(X_0,Y_0)=(x,y)$, for $t\in[0,\tau]$,
300: where $\tau$ is a random time depending on $x$ and $y$. Then we
301: will explain how one can define $(X_t,Y_t)$ for $t\in[0,\infty)$
302: by pasting together different pieces of the trajectory.
303:
304: (i) Recall that the length of any edge is at least $r_0>0$. First
305: suppose that $\d(x,y) \geq 3r_0/4$. Then we let $\{(X_t,Y_t),
306: t\in[0,\tau]\}$ be two independent copies of Brownian motion on
307: $\cS$ and we let $\tau=\inf\{t>0: \d(X_t,Y_t) =r_0/2\}$.
308:
309:
310: (ii) Next suppose that $x,y\in \cS$ are such that $\d(x,y)\in(
311: r_0/4, 3r_0/4)$, and none of these points is a vertex. Let
312: $$ \sigma (r)= {(4|r| -r_0 )^+ \over r_0} \wedge 1,
313: \eqno(3.1)
314: $$
315: and $B$ and $B'$ be independent Brownian motions on $\R$ starting
316: from the origin. Let $U_t=B_t$ and
317: $$dV_t = \sqrt{ 1- \sigma^2 (U_t - V_t)} dB_t +
318: \sigma (U_t - V_t) dB'_t , \eqno(3.2)
319: $$
320: with $V_0=v_0=\d(x,y)> r_0/4$.
321: Then, if we write $Z_t = V_t - U_t$, we obtain
322: $$Z_t = v_0 + \int_0^t (\sqrt{ 1- \sigma^2 (Z_s)}-1) dB_s
323: + \int_0^t \sigma(Z_s) dB'_s,$$
324: and for $Z'_t \df Z_t - r_0/4$,
325: $$Z'_t = v_0 -r_0/4
326: + \int_0^t (\sqrt{ 1- \sigma^2 (Z'_s+r_0/4)}-1) dB_s
327: + \int_0^t \sigma(Z'_s+r_0/4) dB'_s.$$
328: So
329: $$ Z_t'= v_0 -r_0/4 +\int_0^t \gamma (Z_s') dW_s,
330: $$
331: where
332: $$ \gamma(r) \df (\sqrt{ 1- \sigma^2 (r+r_0/4)}-1)^2
333: + \sigma^2(r+r_0/4)
334: $$
335: and $W$ is a Brownian motion on $\R$ with $W_0=0$.
336: The process $Z'$ has the same distribution as
337: $$ t\mapsto v_0 -r_0/4 +W_{\tau_t},
338: $$
339: where
340: $$ \tau_t:= \inf\left\{ s>0: \int_0^s \gamma( v_0 -r_0/4 + W_s)^{-2} ds>t \right\}.
341: $$
342: Note that for small $r>0$, $\gamma (r)=O(r^2)$ and so in particular
343: $\int_{0+} \gamma(r)^{-2} dr = \infty$.
344: Thus by Lemma V.5.2 of [KS],
345: $$ \int_0^{T_0} \gamma (v_0 -r_0/4 + W_s)^{-2} ds =\infty
346: $$
347: almost surely, where $T_0 =\inf\{t>0: v_0 -r_0/4 + W_s=0\}$.
348: We conclude that $Z'$ never hits 0; in other words, $Z$ never reaches
349: $r_0/4$.
350:
351: Suppose that $X_0=x$, $Y_0=y$ and recall that we have assumed that
352: $\d(x,y)\in( r_0/4, 3r_0/4)$. Suppose that $x\in e_1 \in \cE$ and
353: $y\in e_2 \in \cE$. Let $e_1 \setminus \{x\}$ consist of two line
354: segments $e_1^\ell$ and $e_1^r$, with $e_1^r$ being the one closer
355: to $y$. Similarly, $e_2 \setminus \{y\}$ consists of two line
356: segments $e_2^\ell$ and $e_2^r$, and $e_2^\ell$ is closer to $x$.
357: We will define $X$ and $Y$ on an interval $[0,\tau]$ to be
358: specified later.
359: We define $X_t$ on $e_1$ to be such that $\d(X_t, x) = |U_t|$ and
360: $X_t \in e_1^\ell$ if and only if $U_t <0$. We define the process
361: $Y_t$ on $e_2$ by conditions $\d(Y_t, y) = |V_t-v_0|$ and $Y_t
362: \in e_2^\ell$ if and only if $V_t <v_0$.
363: We let $\tau$ be the first time $t>0$ that $X_t$ or $Y_t$ is at a
364: vertex, or $\d(X_t,Y_t) = 3r_0/4$. We see that over the interval
365: $[0, \tau )$, the distance between $X_t$ and $Y_t$ remains in the
366: interval $(r_0/4, \, 3r_0/4)$.
367:
368: \medskip
369:
370: (iii) This part of our argument is based on the ``skew Brownian
371: motion.'' The skew Brownian motion $U$ is a real-valued diffusion
372: which satisfies the stochastic differential equation
373: $$ U_t = B_t + \beta L^U_t, \eqno(3.3) $$
374: where $B$ is a given Brownian motion with $B_0=0$, $\beta \in[-1,1]$
375: is a fixed constant and $L^U$ is the symmetric local time of $U$
376: at $0$, i.e.,
377: $$ L^U_t = \lim_{\eps \to 0} {1\over 2\eps} \int_0^t
378: \bone_{(-\eps, \eps)} (U_s ) ds
379: \, . \eqno(3.4)
380: $$
381: The existence and uniqueness of a strong solution to (3.3)-(3.4)
382: was proved in [HS]. In the special case of $\beta =1$, the
383: solution to (3.3) is the reflected Brownian motion. An alternative
384: way to define the skew Brownian motion is the following. Consider
385: the case $\beta > 0$. Take a standard Brownian motion $B_t'$ and
386: flip every excursion of $B_t'$ from $0$ to the positive side with
387: probability $\beta$, independent of what happens to other
388: excursions (if an excursion is on the positive side, it remains
389: unchanged). The resulting process has the same distribution as $U$
390: defined by (3.3)-(3.4). For more information and references, see
391: recent papers on skew Brownian motion, [BC1] and [BK].
392:
393: Suppose that $x,y\in \cS$, $\d(x,y)\in(r_0/4, 3r_0/4)$, and $x$ is
394: a vertex. Note that $y$ is not a vertex. By assumption, the degree
395: $k$ of vertex $x$ is 3 or greater. Suppose that $B_t$ is a
396: Brownian motion on $\R$ and let $U$ be a solution to (3.3)-(3.4),
397: with $\beta$ defined by $(1-\beta)/(1+\beta) = k-1$. Note that
398: $\beta<0$. We label negative excursions of $U$ from 0 with numbers
399: $1, 2, \dots, k-1$, in such a way that every excursion has a label
400: chosen uniformly from this set and independently of all other
401: labels.
402:
403: Suppose that $B'$ is a Brownian motion independent of $B$. Recall
404: the definition of the function $\sigma$ and the process $V$ given
405: in (3.1) and (3.2), respectively, with $V_0=v_0=\d(x,y)$. Then,
406: if we write $Z_t = V_t - U_t$, we obtain
407: $$Z_t = v_0 -\beta L^U_t
408: +\int_0^t (\sqrt{ 1- \sigma^2 (Z_s)}-1) dB_s
409: + \int_0^t \sigma(Z_s) dB'_s,$$
410: and for $Z'_t \df Z_t - r_0/4$,
411: $$Z'_t = v_0 -r_0/4 -\beta L^U_t
412: + \int_0^t (\sqrt{ 1- \sigma^2 (Z'_s+r_0/4)}-1) dB_s
413: + \int_0^t \sigma(Z'_s+r_0/4) dB'_s.$$
414: We have already pointed out that for small $r>0$,
415: $$ \gamma(r) \df (\sqrt{ 1- \sigma^2 (r+r_0/4)}-1)^2
416: + \sigma^2(r+r_0/4) =O(r^2).
417: $$
418: Since $\beta<0$, the process $- \beta L^U_t$ is nondecreasing.
419: These observations and the argument used in the first half of (ii)
420: imply that $Z'$ never hits 0, i.e., $Z$ never reaches $r_0/4$.
421:
422: The ball $\cB(x,3r_0/4)$ consists of line segments $I_j$,
423: $j=1,2,\dots,k$. We assume that $I_k$ is the line segment
424: containing $y$. We define $X_t$ on these line segments so that
425: $\d(X_t, x) = |U_t|$. If $U_t >0$ then $X_t \in I_k$. If $U_t < 0$
426: then $X_t \in I_j$, where $j$ is the label of the excursion of $U$
427: straddling $t$. Suppose that $y\in e\in \cE$. Let $e \setminus
428: \{y\}$ consist of two line segments $e^\ell$ and $e^r$, with
429: $e^\ell$ being the one closer to $x$. We define $Y_t$ on $e$ by
430: $\d(Y_t,y)=|V_t-v_0|$.
431: We let $Y_t \in e^\ell$ if and only if $V_t <v_0$. We let $\tau$ be the
432: infimum of $t$ such that $Y_t$ is at a vertex, or $\d(X_t,Y_t) =
433: 3r_0/4$. Observe that over the interval $[0, \tau )$, the distance
434: between $X_t$ and $Y_t$ remains in the interval $(r_0/4, \,
435: 3r_0/4)$.
436:
437: \medskip
438:
439: Now we will define the process $(X_t,Y_t)$ for all $t\geq 0$,
440: assuming that $X_0 =x$, $Y_0=y$ and $\d(x,y) > r_0/4$. We use one
441: of the parts (i)-(iii) of the proof to define the process $(X,Y)$
442: on an interval $[0,\tau_1]$. Then we proceed by induction. Suppose
443: that the process has been defined on an interval $[0,\tau_k]$ and
444: $\d(X_{\tau_k},Y_{\tau_k})>r_0/4$. Then we use the appropriate
445: part (i)-(iii) of the proof to extend the process, using the
446: strong Markov property at $\tau_k$, to an interval
447: $[0,\tau_{k+1}]$. It is easy to see that $\tau_k\to\infty$ a.s.,
448: so the process $(X_t,Y_t)$ is defined for all $t\geq 0$. It is
449: straightforward to check that $\{X_t, t\geq 0\}$ and $\{Y_t, t\geq
450: 0\}$ are Brownian motions on $\cS$ and $(X,Y)$ is a shy Markov
451: coupling, as defined in Section 2.
452: \qed
453:
454: \bigskip
455: \noindent{\bf Remark 3.2}. One may wonder whether it is possible
456: to construct the shy coupling in the proof of Theorem 3.1 using
457: the skew Brownian motion in such a way that the distance between
458: $X$ and $Y$ does not change on time intervals where both processes
459: stay away from vertices (``synchronous coupling''). This idea
460: works well for many graphs but runs into technical problems when
461: we have a configuration similar to that in
462: Fig.~3.1, with many
463: geodesics joining two vertices. Suppose that the lengths of edges in
464: Fig.~3.1
465: are chosen so that there are 6 geodesics between $x$
466: and $y$. We will argue that if $X_0=x$ and $Y_0=y$ then for small
467: $t>0$ the distance between $X$ and $Y$ has to decrease. This is
468: because $X$ will have to move towards $z$ with probability $5/7$,
469: and $Y$ will have to move towards $z$ with probability $1/3$.
470: Since $5/7 + 1/3>1$, $X$ and $Y$ will find themselves on a
471: geodesic from $x$ to $y$, moving away from their starting points
472: towards each other, with positive probability. For this reason, we
473: could not find a ``synchronous'' shy coupling based on the skew
474: Brownian motion.
475:
476: \bigskip \vbox{ \epsfxsize=3.0in
477: \centerline{\epsffile{shy3.eps}}
478:
479: \centerline{Figure 3.1.} }
480: \bigskip
481:
482: The rest of this section is devoted to graphs that have at least
483: one vertex of degree 1 (recall that vertices of degree 2 can be
484: ignored and we assume that $\cG$ does not contain any of them). We
485: do not have a general theorem covering all graphs with some
486: vertices of degree 1 but we have four examples illustrating some
487: special cases.
488:
489: \bigskip
490: \noindent{\bf Example 3.3}. This example is similar to Examples
491: 2.3 (i)-(iii). Suppose that there exists an isometry $I: \cS\to
492: \cS$ such that $\inf_{x\in\cS} \d(x, I(x))>0$. It is not hard to
493: show that this holds if the isometry has no fixed points, i.e., if
494: there does not exist $x\in \cS$ with $I(x) =x$. If such an
495: isometry exists, then we can first construct the process $X$ and
496: then take $Y_t = I(X_t)$ for all $t\geq 0$. Obviously, thus
497: constructed coupling $(X,Y)$ is shy. Fig.~3.2 shows that a graph
498: with some vertices of degree 1 may have this property.
499:
500: \bigskip \vbox{ \epsfxsize=1.5in
501: \centerline{\epsffile{shy4.eps}}
502:
503: \centerline{Figure 3.2.} }
504: \bigskip
505:
506: Our next lemma is a large deviations-type estimate. Recall that
507: all edges are at least $r_0>0$ units long, by assumption.
508:
509: \bigskip
510:
511: Let $X$ be Brownian motion on a graph $\cS$. For $A\subset \cS$, define
512: $T_A = \inf\{t\geq 0: X_t\in A\}$.
513:
514: \bigskip
515:
516: \noindent{\bf Lemma 3.4}. {\sl Assume that the degrees of all
517: vertices are bounded above by $m_0$. There exist constants
518: $c_0>0$, $t_0<\infty$ depending on $r_0$ only such that for
519: $t\in(0,t_0)$ and $r>0$ with $r^2> t$ and $\cB(x,r)^c \ne
520: \emptyset$, we have
521: $$
522: \left( {c_0\over m_0} \sqrt{t \over 2 \pi} {1\over 2 r}
523: \right)^{[r/r_0]} \exp\left(-{r^2\over 2t}\right)
524: \leq \P(T_{\cB(x,r)^c} < t) \leq
525: (m_0^{[r/r_0]})!
526: \sqrt{2t \over \pi}
527: {1\over r} \exp\left(-{r^2\over 2t}\right).$$
528:
529: }
530:
531: \bigskip
532:
533: In applications of the above estimates, $r$ will be ``fixed'' and
534: then a $t$ (much smaller than $r^2$) will be chosen. For this
535: reason, we did not try to optimize the non-exponential
536: factors---most likely they are not best possible.
537:
538: \bigskip
539: \noindent{\bf Proof}. (i) First we will prove the lower bound. We
540: start with some preliminary estimates.
541:
542: Suppose that $B$ is a Brownian motion on $\R$ with $B_0=0$ and let
543: $T^B_r=\inf\{t\geq 0: B_t =r\}$. Then $\P(T^B_r < t) = 2 {1\over
544: \sqrt{2 \pi t}} \int_r^\infty e^{-u^2/2t} du$ for $r>0$. The
545: following inequalities are well known (see Problem 9.22 on page
546: 112 of [KS]):
547: $${r \over 1 + r^2} e^{-r^2/2} \leq
548: \int_r^\infty e^{-u^2/2} du \leq {1 \over r} e^{-r^2/2}.
549: $$
550: So for $r\geq 1$,
551: $$ {1 \over 2 r} e^{-r^2/2} \leq
552: \int_r^\infty e^{-u^2/2} du \leq {1 \over r} e^{-r^2/2}.
553: $$
554: By scaling we obtain for $t\leq r^2$,
555: $$ \sqrt{t \over 2 \pi}
556: {1\over r} \exp\left(-{r^2\over 2t}\right)
557: \leq \P (T^B_r<t)= \sqrt{{2\over \pi}} \int_{r/\sqrt{t}}^\infty
558: e^{-v^2/2} dv
559: \leq \sqrt{2t \over \pi}
560: {1\over r} \exp\left(-{r^2\over 2t}\right).\eqno(3.5)$$
561: For $r\leq 1$, we have a trivial lower bound $ \int_1^\infty
562: e^{-u^2/2} du \df c_0>0$. For $t\geq r^2$, the same upper bound
563: holds but the lower bound has to be replaced with a trivial bound
564: $c_1=\sqrt{2\over \pi} c_0>0$.
565:
566: Let $t_0< \infty$ be the largest real such that
567: $$ 1 - 2 \exp(- r_0^2/ (2t_0))\geq 1/2 .
568: $$
569: We will derive an estimate for $\P(T^B_r < t\land T^B_{-r_0/2})$,
570: for $r\geq r_0/2$ and $t< t_0$. If $r^2<t_0$, then $r\leq c_2 r_0$
571: for some constant $c_2 < \infty$. In this case, it follows easily
572: from the support theorem for Brownian motion that $\P(T^B_r <
573: t\land T^B_{-r_0/2})> c_3>0$ for every $t\in (r^2, \, t_0)$.
574:
575: Now suppose that $t\leq r^2\land t_0$. If Brownian motion hits
576: $-r_0/2$ and then it reaches $r$ in $t$ seconds or less, it has to
577: go from level $-r_0/2$ to level $r$ in $t$ seconds or less. Hence,
578: by the strong Markov property applied at $T^B_{-r_0/2}$,
579: $$\eqalignno{
580: \P&(T^B_r < t\land T^B_{-r_0/2}) \geq
581: \P(T^B_r < t) - \P(T^B_{r+r_0/2} < t)\cr
582: &\geq \sqrt{t \over 2\pi}
583: {1\over r} \exp\left(-{r^2\over 2t}\right)
584: - \sqrt{2t \over \pi}
585: {1\over r+r_0/2} \exp\left(-{(r+r_0/2)^2\over 2t}\right)\cr
586: &\geq \sqrt{t \over 2 \pi}
587: {1\over r} \exp\left(-{r^2\over 2t}\right)
588: - 2\sqrt{t \over 2 \pi}
589: {1\over r} \exp\left(-{r^2\over 2t}\right)
590: \exp\left(-{ r_0^2\over 2t}\right)\cr
591: & = \sqrt{t \over 2 \pi} {1\over r}
592: \exp\left(-{r^2\over 2t}\right)
593: \left[ 1 - 2 \exp\left(-{ r_0^2\over 2t}\right)\right]\cr
594: &\geq \sqrt{t \over 2 \pi} {1\over 2 r}
595: \exp\left(-{r^2\over 2t}\right).&(3.6)
596: }$$
597:
598: \medskip
599:
600: Suppose that $y_1$ is a vertex, $\d(y_1,y_2) = r_0/2$, and $y_3$
601: lies between $y_1$ and $y_2$ so that $\d(y_1,y_3) + \d(y_3, y_2) =
602: r_0/2$. Let $\d(y_1,y_3) =r_1 \in[0,r_0/2]$ and suppose that $X_0
603: = y_3$. The process $R_t \df \d(X_t, y_1)$ is a one-dimensional
604: reflected Brownian motion with $R_0=r_1$, at least until it
605: reaches $r_0$. Let $\{z_1, z_2, \dots, z_k\}$ be the set of all
606: points with $\d(z_j, y_1) = r_0/2$ ($y_2$ is one of these points).
607: Let $T^R_{r_0/2} = \inf\{t\geq 0: R_t = r_0/2\}$. If $X$ visited
608: $y_1$ before $T^R_{r_0/2}$ then it is at $y_2$ at time
609: $T^R_{r_0/2}$ with probability $1/k$, by symmetry. Since $k\leq
610: m_0$, the probability that $X$ starts at $y_3$ and reaches $y_2$
611: in $s$ seconds or less is greater than or equal to the probability
612: that reflected Brownian motion that starts from $r_1$ reaches
613: $r_0/2$ in $s$ seconds or less, divided by $m_0$. The last
614: probability is bounded below by the analogous probability for the
615: non-reflected Brownian motion, so using (3.5) we obtain,
616: $$\P(T^X_{y_2} \leq s \mid X_0 = y_3) \geq
617: {1\over m_0} \sqrt{s \over 2\pi}
618: {1 \over r_0/2-r_1}
619: \exp\left(-{(r_0/2 - r_1)^2\over 2s}\right), \eqno(3.7)
620: $$
621: for $s \leq (r_0/2 - r_1)^2$. If $s \geq (r_0/2 - r_1)^2$, the
622: bound is $c_1/m_0$. Note that these estimates hold for all $r_1
623: \in[0,r_0/2]$, including $r_1 =0$.
624:
625: Consider any $x'\in \prt \cB(x,r)$ and let $\Gamma\subset \cS$ be
626: a geodesic connecting $x$ and $x'$. Suppose that $\Gamma $
627: contains some vertices and denote them $x_1, x_2, \dots, x_k$, in
628: order in which they lie on $\Gamma$, going from $x$ to $x'$. Let
629: $x_0 = x$ and $x_{k+1} = x'$. If there is a vertex closer to $x_0$
630: than $r_0/2$ and it is not $x_1$ then we let $y_0$ be the point at
631: the distance $r_0/2$ from that vertex, between $x_0$ and $x_1$.
632: For every $x_j$, $j\geq 1$, we let $y_j\in \Gamma$ be the point
633: $r_0/2$ away from $x_j$, between $x_j$ and $x_{j+1}$.
634:
635: Let $z_j$, $j=1, \dots, m_1$, be the sequence of all points $x_j$
636: and $y_j$, in the order in which they appear on $\Gamma$ from $x$
637: to $x'$, including $x$ and $x'$. Note that $\sum_{1\leq j \leq
638: m_1-1}\d(z_j, z_{j+1})=r$. By the strong Markov property applied
639: at the hitting times of $z_j$'s, the probability that $X$ starting
640: from $x$ will hit $x'$ in $t$ seconds or less is bounded below by
641: $\prod_{j=1}^{m_1-1} p_j$, where $p_j$ is the probability that $X$
642: starting from $z_j$ will hit $z_{j+1}$ in $t_j$ seconds or less,
643: and $t_j = t \d(z_j, z_{j+1})/r$.
644:
645: If $z_j$ is a vertex or $x_0$ then, by (3.7), if $t_j \leq \d(z_j,
646: z_{j+1})^2$,
647: $$p_j \geq
648: {1\over m_0} \sqrt{t_j \over 2\pi}
649: {1 \over \d(z_j, z_{j+1})}
650: \exp\left(-{\d(z_j, z_{j+1})^2\over 2t_j}\right).\eqno(3.8)$$
651: If $t_j \geq \d(z_j, z_{j+1})^2$ then
652: $$p_j \geq c_1/m_0 \geq (c_1/m_0)
653: \exp\left(-{\d(z_j, z_{j+1})^2\over 2t_j}\right).\eqno(3.9)$$
654: For other $z_j$'s we use (3.6) to obtain, for $t_j \leq \d(z_j,
655: z_{j+1})^2$,
656: $$p_j \geq
657: \sqrt{t_j \over 2\pi}
658: {1 \over 2\d(z_j, z_{j+1})}
659: \exp\left(-{\d(z_j, z_{j+1})^2\over 2t_j}\right),\eqno(3.10)$$
660: and for $t_j \geq \d(z_j, z_{j+1})^2$,
661: $$p_j \geq c_3 \geq c_3
662: \exp\left(-{\d(z_j, z_{j+1})^2\over 2t_j}\right).\eqno(3.11)$$
663:
664: The product of exponential factors on the right hand sides of
665: (3.8)-(3.11) is equal to
666: $$ \prod_{j=1}^{m_1-1} \exp \left( - {\d(z_j,z_{j+1})^2
667: \over 2 t_j}\right )
668: = \exp \left( - {r^2 \over 2 t}\right ).\eqno(3.12)
669: $$
670:
671: If $t_j \geq \d(z_j,z_{j+1})^2 $ then the non-exponential factor
672: in (3.9) is $c_1/m_0$ and it is $c_3$ in (3.11). The
673: non-exponential factors in (3.8) and (3.10) are bounded below by
674: $${1\over m_0} \sqrt{t_j \over 2 \pi} {1\over 2 \d(z_j,z_{j+1})}
675: = {1\over m_0} \sqrt{t \d(z_j, z_{j+1})/r \over 2 \pi}
676: {1\over 2 \d(z_j,z_{j+1})}
677: \geq {1\over m_0} \sqrt{t \over 2 \pi}
678: {1\over 2 r}.$$
679: We conclude that the product of non-exponential factors in
680: (3.8)-(3.11) is bounded below by
681: $$\left( {c_4\over m_0} \sqrt{t \over 2 \pi} {1\over 2 r}
682: \right)^{m_1-1}
683: \geq \left( {c_4\over m_0} \sqrt{t \over 2 \pi} {1\over 2 r}
684: \right)^{[r/r_0]}.
685: $$
686: This
687: combined
688: with (3.12) gives the lower bound in the lemma.
689:
690: (ii) Next we will prove the upper bound. Let $\{\Gamma_j\}$ be the
691: family of all Jordan arcs in $\cS$ linking $x$ with $\prt
692: \cB(x,r)$. The number of edges in $\cB(x,r)$ is bounded by $m_2 =
693: m_0^ {[r/r_0]}$ so the number of $\Gamma_j$'s is bounded by $m_3=
694: m_2!$. The length of any $\Gamma_j$ is at least $r$.
695:
696: Consider some $\Gamma_k$. We will define a process $R^k_t$ that
697: measures the distance from $X_t$ to $x$ along $\Gamma_k$, in a
698: sense. We will ``erase'' excursions away from $\Gamma_k$ and loops
699: as follows. For $t>0$, let $\ell(t) = \sup\{s\leq t: X_s \in
700: \Gamma_k\}$. Let $\cV_k= \cV\setminus \Gamma_k$, i.e., $\cV_k$ is
701: the set of vertices that do not belong to $\Gamma_k$. Let $T_1=0$,
702: $$\eqalign{
703: S_j &= \inf\{t\geq T_j: X_t \in \cV_k\}, \quad j \geq 1,\cr
704: T_{j+1} & = \inf\{t\geq S_{j}: X_t = X_{\ell(S_{j})}\}, \quad j\geq 1.
705: }$$
706: If $t\in [S_j, T_{j+1}]$ for some $j\geq 1$, we let $R^k_t$ be the
707: distance from
708: $X_{\ell(S_{j})}$
709: to $x$ along $ \Gamma_k$. For other $t$, we let $R^k_t$ be the
710: distance from $X_{\ell(t)}$ to $x$ along $\Gamma_k$. The process
711: $R^k_t$ is a time-change of reflected Brownian motion, that is, it
712: is reflected Brownian motion ``frozen'' on time intervals when $X$
713: is outside $\Gamma_k$ (and some other intervals). Hence, the
714: probability that $R^k_t$ reaches $r$ in $t$ seconds or less is
715: less than the right hand side of (3.5). Note that one of the
716: processes $R^k_t$ must be at the level $r$ at the time when $X$
717: hits $\prt \cB(x,r)$. Hence, an upper bound on the probability in
718: the statement of the lemma is the product of the right hand side
719: of (3.5) and $m_3$.
720: \qed
721:
722: \bigskip
723: \noindent{\bf Example 3.5}. Suppose that the graph $\cS$ is
724: compact and has the following structure. For some $x\in \cS$, the
725: set $\cS\setminus \{x\}$ is disconnected and consists of a finite
726: number of disjoint finite trees $\cT_1, \cT_2, \dots, \cT_k$, and
727: a graph $\cU$ (not necessarily a tree). We say that a vertex of a
728: tree is a leaf if it has degree 1. Assume that for some
729: $r_1>r_2>0$ and every leaf $y\ne x$ of any tree $\cT_j$ we have
730: $\d(x,y) \geq r_1$, and for any $z\in \cU$, $\d(x,z) \leq r_2$
731: (see, for example, Fig.~3.3). Suppose that $(X,Y)$ is a coupling
732: of Brownian motions on $\cS$. We will show that
733: $(X,Y)$ is not a shy coupling.
734:
735: \bigskip \vbox{ \epsfxsize=2.5in
736: \centerline{\epsffile{shy6.eps}}
737:
738: \centerline{Figure 3.3.} }
739: \bigskip
740:
741: (i) Let $\wt \cT = \bigcup _j \cT_j \setminus \{x\}$. First, we
742: will show that there exist $p_1>0$ and a stopping time
743: $T_1<\infty$ such that with probability greater than $p_1$, either
744: $X_{T_1}=x$ and $Y_{T_1}\in \wt \cT$, or $Y_{T_1}=x$ and
745: $X_{T_1}\in \wt \cT$. Let $r_3 \in (r_2, r_1)$ and $\cW =
746: \{y\in\wt \cT: \d(y,x)\geq r_3\}$. It is easy to see that $X$ is
747: recurrent so $T_2 \df \inf\{t\geq 0: X_t \in \cW\} < \infty$ a.s.
748: Suppose first that $Y_{T_2} \in \wt \cT\cup \{x\}$ and let $T_3
749: =\inf\{t\geq T_2: X_t = x \hbox { or } Y_t =x\}$. Then $T_1 =
750: T_3$ has the properties stated above.
751:
752: Next suppose that $Y_{T_2} \in \cU$ and let $T_4 = \inf\{t\geq T_2:
753: X_t = x\}$ and $T_5 = \inf\{t\geq T_2: Y_t = x\}$. By Lemma 3.4,
754: for some $p_2, t_0>0$ and all $y\in \cU$ and $z\in \cW$,
755: $$
756: \P(T_5< T_2 + t_0\mid X_{T_2}=z, Y_{T_2}=y)
757: > \P(T_4 < T_2 + t_0\mid X_{T_2}=z, Y_{T_2}=y) + p_2.
758: $$
759: Hence, $\P(T_5 < T_4\mid X_{T_2}=z, Y_{T_2}=y) > p_2$, and it follows
760: that we can take $T_1 = T_5$ on the event $\{T_5 < T_4\}$. This
761: completes the proof of our claim, with $p_1=p_2$.
762:
763:
764: (ii)
765: Recall that the length of any edge is bounded below by $r_0>0$.
766: Fix an arbitrarily small $\eps\in(0, r_0/3)$. We will show in the
767: remaining part of the proof that $X$ and $Y$ come within $\eps$
768: distance to each other in finite time almost surely, which will
769: then imply that $(X,Y)$ is not a shy coupling.
770:
771: The rest of the proof is based on an inductive argument. We will
772: now formulate and prove the inductive step.
773:
774:
775: Suppose that for some $x_0\in \cS$, $\cS \setminus\{x_0\}=\cU_1
776: \cup \cU_2$, where $\cU_1$ and $\cU_2$ are disjoint and $\cU_1$ is
777: a finite union of finite trees $\cW_j$, $j=1,2,\dots, k$. Assume
778: that $Y_0 = x_0$ and $X_0 \in \cU_1$ (the argument is analogous
779: if the roles of $X$ and $Y$ are interchanged). Suppose without
780: loss of generality that $X_0 \in \cW_1$. Let $x_1\ne x_0$ be the
781: vertex of $\cW_1$ closest to $x_0$, and $\cW_1 \setminus \{x_1\} =
782: \cW_2 \cup \cW_3$, where $\cW_2$ and $\cW_3$ are disjoint, and
783: $\cW_3$ is the edge joining $x_0$ and $x_1$. We will first assume
784: that $\cW_2\ne \emptyset$. We will show that for some $p_3>0$ and
785: some stopping time $T_6< \infty$, with probability greater than
786: $p_3$, we either have $\d(X_{T_6}, Y_{T_6})\leq \eps$ or $Y_{T_6} =
787: x_1$ and $X_{T_6} \in \cW_2$.
788:
789: If $\d(X_0, Y_0)\leq \eps$ then we can take $T_6 = 0$.
790:
791: Assume that $\d(X_0, Y_0)> \eps$. Let $x_2\in \cW_3$ be the point
792: with $\d(x_0,x_2) = \eps/3$. Note that $\d(x_2,Y_0) \leq (1/2)
793: \d(x_2, X_0)$. Let $T_7 = \inf\{t\geq 0: X_t = x_2\}$ and $T_8 =
794: \inf\{t\geq 0: Y_t = x_2\}$. By Lemma 3.4, for some $s>0$,
795: $$\P(T_8 < s) > \P(T_7 < s).$$
796: Hence, with probability $p_3>0$, $T_8< T_7$ and either $X$ and $Y$
797: have met by the time $T_8$, or $X$ is on the opposite side of
798: $Y_{T_8}$ in $\cS$ than $x_0$. Let $T_{9} = \inf\{t\geq T_8: Y_t =
799: x_1\}$ and $T_{10} = \inf\{t\geq T_8: Y_t = x_0\}$. Since $\cS$
800: contains only a finite number of finite trees $\cT_k$, there is an
801: upper bound on the edge length in any tree $\cT_k$, say,
802: $\rho<\infty$. This and the fact that $\d(x_2, x_0) =\eps/3$ imply
803: that $\P(T_9 < T_{10}) \geq p_4$ for some $p_4>0$ that may depend
804: on $\eps$. If the events $\{T_8< T_7\}$ and $\{T_9 < T_{10}\}$
805: hold then either $X$ and $Y$ have met by the time $T_9$ or
806: $Y_{T_9})= x_1$ and $X_{T_9})\in \cW_2$.
807:
808: We note that if $\cW_2=\emptyset$ (i.e., $\cW_1 $ is a single
809: edge) then the same argument proves that for some $p_3>0$ and some
810: stopping time $T_6< \infty$, we have $\d(X_{T_6}, Y_{T_6})\leq \eps$
811: with probability greater than $p_3p_4$.
812:
813: (iii) Let us rephrase the claim proved in step (ii). We have shown
814: that for some $p_5\df p_3p_4>0$ and some stopping time $T_6<
815: \infty$, with probability greater than $p_5$, we either have
816: $\d(X_{T_6}, Y_{T_6})\leq \eps$ or $Y_{T_6}$ and $X_{T_6}$ satisfy
817: the same assumptions as $Y_0$ and $X_0$, but relative to graphs
818: $\wt \cU_1 \df \cW_2$ and $\wt \cU_2 \df \cS\setminus (\{x_2\}
819: \cup \cW_2)$ in place of $\cU_1$ and $\cU_2$. Recall the claim
820: proved in part (i) of the proof and the final remark in step (ii).
821: Note that $\wt \cU_1$ has at least one edge less than $\cU_1$ so
822: by induction, we can repeat the inductive step (ii) a finite
823: number of times and show that with a probability $p_6>0$, $X$ and
824: $Y$ come within $\eps$ of each other before some time $t_1<
825: \infty$. It is easy to check that $p_6$ and $t_1$ can be chosen so
826: that they do not depend on the starting points of $X$ and $Y$. The
827: Markov property and induction can be used to show that $X$ and $Y$
828: have to come within $\eps$ of each other by the time $jt_1$ with
829: probability greater than $1-(1-p_6)^j$. We let $j\to\infty$ to see
830: that $X$ and $Y$ come within $\eps$ of each other at some finite
831: time with probability one. Since $\eps \in (0, r_0/3)$ is
832: arbitrary, the coupling is not shy.
833: \qed
834:
835: \bigskip
836: \noindent{\bf Example 3.6}. Suppose that $\cS$ is a tree with the
837: property that it has a ``backbone'' that is topologically a line,
838: with a finite or countable number of finite trees attached to it.
839: See Fig.~3.4 for an example.
840: Recall that we have assumed that each edge has length at least
841: $r_0>0$.
842:
843: \bigskip \vbox{ \epsfxsize=3.0in
844: \centerline{\epsffile{shy7.eps}}
845:
846: \centerline{Figure 3.4.} }
847: \bigskip
848:
849: Let $\dots, x_{-2}, x_{-1}, x_0, x_1, x_2, \dots$ be the sequence
850: of points along the ``backbone'' $\cU$ where the side trees are
851: attached (the sequence can be finite or it can extend to infinity
852: in one or two directions). If the sequence extends to infinity in
853: both directions and
854: the graph is invariant under a non-constant shift,
855: then, according to Example 3.3, there exists a
856: shy coupling.
857:
858: Assume that
859: \item{(i)} the diameters of the side trees are
860: uniformly bounded and
861: \item{(ii)} $\{x_k\}$ does not extend to
862: infinity in both directions, or $\{x_k\}$ extends to infinity in
863: both directions but
864: the family $\{x_k\}$ is not shift-invariant, i.e., for every $c\ne
865: 0$, there exists $x_j$ such that $x_j + c \ne x_k$ for all $k$.
866:
867: We will show that under these assumptions there is no shy
868: coupling.
869:
870: Let $Q$ be the ``projection'' of $\cS$ on $\cU$, i.e., $Q(x) = x$
871: for $x\in \cU$, and $Q(x) = x_k$ if $x$ belongs to a tree which is
872: attached to $\cU$ at $x_k$. We will identify the ``backbone''
873: $\cU$ with the real line so that we can think of $Q(X)$ and $Q(Y)$
874: as real-valued processes.
875: Let $X$ and $Y$ be a coupling of Brownian motions on $\cS$.
876: If $X$ makes an excursion into a side tree then $Q(X)$ remains
877: constant on the excursion interval (including the endpoints).
878: Hence, $Q(X)$ and $Q(Y)$ are continuous processes. It is easy to
879: see that they are local martingales. Informally speaking, they are
880: Brownian motions frozen on some random intervals. The process $Z_t
881: = Q(X_t) - Q(Y_t)$ is also a continuous local martingale. Suppose
882: without loss of generality that $Z_0 >0$ and let
883: $$ T_0 =\inf\{t\geq 0: Z_t =0\} .
884: $$
885: Then $Z_{T_0\land t}$ is a non-negative local martingale and so it must
886: have an almost surely finite limit $Z_\infty$ on $\{T_0=\infty\}$.
887:
888: We will show first that $\P(T_0=\infty)=\P(T_0=\infty \hbox{ and }
889: Z_\infty =0)$, in other words, $\P(T_0=\infty \hbox{ and }
890: Z_\infty >0)=0$. Let $B$ be a Brownian motion on $\R$ with
891: $B_0=0$. Consider a small $\eps>0$ and $t_0>0$, let $\delta \in
892: (0, \, \eps)$ be such that
893: $$
894: \P \left(\sup_{t\in[0,t_0]} |B_t| < \delta \right) < p_1/2,
895: \qquad \hbox{where } \ \
896: p_1:= \P \left(\sup_{t\in [0,t_0]}|B_{t} | <\eps \right).
897: $$
898:
899: Let $T_1$ be the first time when all of the following conditions
900: hold: $\d(X_{T_1},\cU) \geq \eps$, the distance from $X_{T_1}$ to
901: any vertex of $\cS$ is greater than $\eps$, $Y_{T_1} \in \cU$, and
902: $\inf_k \d(Y_{T_1}, x_k) \geq \delta$ (the argument is analogous
903: if the roles of $X$ and $Y$ are interchanged). If $T_1<\infty$
904: then with probability $p_1/2$ or greater, $X$ will stay on the
905: same side tree over the interval $[T_1, T_1+t_0]$, while $Y$ will
906: move away from $Y_{T_1}$ by more than $\delta$ units over the same
907: time interval. Hence with probability $p_1/2$ or greater, $Z_t$
908: will have an oscillation of size at least $\delta$ over the
909: interval $[T_1, T_1+t_0]$. We proceed by induction. If $T_k <
910: \infty$ then we define $T_{k+1}=T_1 \circ \theta_{T_k+t_0} +
911: T_k+t_0$, where $\theta_\cdot$ is the usual Markovian shift
912: operator. Then with probability greater than $p_1/2$, $Z_t$ has an
913: oscillation of size at least $\delta$ over the interval $[T_{k+1},
914: T_{k+1}+t_0]$, independent of whether that happened over any
915: interval $[T_j, T_j+t_0]$, $j\leq k$. Hence, with probability one,
916: either $T_k= \infty$ for some $k$ or $Z_t$ has an infinite number
917: of oscillations of size $\delta$ over disjoint intervals of length
918: $t_0$, and, therefore in the latter case, $Z_t$ does not have a
919: limit as $t\to \infty$. Applying the above argument to a
920: decreasing sequence of $\{\eps_n, n\geq 1\}$ and a decreasing
921: sequence of $\{\delta_n, n \geq 1\}$ both tending to zero and
922: after deleting a null set from $\Omega$, we may and do assume that
923: for every $\omega \in \Omega$ and for every $\eps_n$, there is
924: some $N>1$
925: such that for every $j\geq N$, with $\eps_n$ and $\delta_j$ in place of
926: $\eps$ and $\delta$ above, either $T_k (\omega) =\infty$ for some
927: $k$ or $Z_t (\omega) $ does not have a limit as $t\to \infty$. The
928: processes $X$ and $Y$ are recurrent because the one-dimensional
929: Brownian motion is. Hence,
930: for every $x_j$,
931: each one of them will enter the side tree attached to $\cU$ at
932: $x_j$ infinitely often. After deleting a null set from $\Omega$,
933: we may and do assume that the aforementioned property holds for
934: every $\omega\in \Omega$.
935:
936: For $\omega \in \{T_0=\infty \hbox{ and } Z_\infty
937: >0 \}$, let
938: $$ c(\omega) =\lim_{t\to\infty} Z_t (\omega)>0 .
939: $$
940: We choose an $x_j$, relative to $c(\omega)$, as follows. If $\{x_n
941: \}$ does not extend to $-\infty$ ($\infty$) then we let $x_j$ be
942: the leftmost (rightmost, resp.) point of the sequence. Otherwise
943: we fix an $x_j$ with the property that $x_j + c (\omega) \ne x_k$
944: for all $k$ (such an $x_j$ exists by assumption).
945: Note that both $X(\omega)$ and $Y(\omega)$
946: enter the side tree attached to $\cU$ at $x_j$ infinitely often.
947: Hence one can find some $\eps>0$ from $\{\eps_n, n\geq 1\}$ and
948: $\delta>0$ from $\{\delta_n, n\geq 1\}$, and an increasing
949: sequence of random times $\{S_k, k\geq 1\}$ with $\lim_{k\to
950: \infty}S_k=\infty$ such that all of the following hold. One has
951: $\d(X_{S_k},\cU) \geq \eps$, the distance from $X_{S_k}$ to any
952: vertex of $\cS$ is greater than $\eps$, $Y_{S_k} \in \cU$, and
953: $\inf_n \d(Y_{S_k}, x_n) \geq \delta$ for every $k\geq 1$ (or the
954: statement will hold with the roles of $X$ and $Y$ interchanged).
955:
956: Hence, all stopping times $\{T_k, k\geq 1\}$ defined in the
957: proceeding paragraph are finite. We have shown that this event
958: implies that $Z_\infty (\omega )$ does not exist. This
959: contradiction proves that $\P(T_0=\infty \hbox{ and }
960: Z_\infty>0)=0$ and therefore $\P(T_0=\infty)=\P(T_0=\infty \hbox{
961: and } Z_\infty =0)$.
962:
963: Recall that we have assumed that all the edges have length at
964: least $r_0>0$. So on $\{T_0=\infty \hbox{ and } Z_\infty =0\}$, by
965: the recurrence of the one-dimensional Brownian motion, we have
966: $\liminf_{t\to \infty}\d(X_t, Y_t)=0$. We now only need to exam
967: $\omega \in \{T_0<\infty\}$ and to prove $X(\omega)$ and
968: $Y(\omega)$ will come arbitrarily close to each other.
969:
970: Consider any $\eps\in (0, r_0/4)$, where $r_0>0$ is a lower bound
971: for the length of any edge in $\cS$. We want to show that with
972: probability one, there exists $t$ such that $\d(X_t, Y_t)\leq
973: \eps$. We have already proved that $\liminf_{t\to \infty} \d(X_t,
974: Y_t)=0$ on $\{T_0=\infty\}$. On $\{T_0<\infty\}$, at time $T_0$,
975: either $X_{T_0}=Y_{T_0}$ or one of processes $\{X_{T_0},
976: Y_{T_0}\}$ is at some $x_k$ and the other process is in a side
977: tree $\cT$ attached to $\cU$ at $x_k$. Without loss of generality,
978: assume that $X_{T_0}=x_k$ and $Y_{T_0}\in \cT$. If $\d(X_{T_0},
979: Y_{T_0})\leq\eps$ then we are done. Suppose that $\d(X_{T_0},
980: Y_{T_0})>\eps$. Let $z_0$ be the point at the edge $e$ of $\cT$
981: that is attached to $\cU$ with $\d(z_0, x_k)=\d(z_0, \cU)=\eps/4$.
982: Let $S_1$ be the first time after $T_0$ when $X_t = z_0$. By Lemma
983: 3.4, with probability $p_2>0$, $X_t$ reaches $z_0$ after $T_0$
984: before $Y_t$ gets there. If this event occurs, both $X_{S_1}$ and
985: $Y_{S_1}$ will have distance at least $\eps/4$ away from $\cU$.
986: Let $R_1$ be the first time after $S_1$ when both $X_t$ and $Y_t$
987: are outside $\cT$. An argument analogous to that in parts (ii) and
988: (iii) of Example 3.5 shows that with probability $p_3>0$, the
989: processes $X$ and $Y$ will meet during the time interval $[S_1,
990: R_1]$. In other words, conditioning on $\{T_0<\infty\}$ and
991: $\d(X_{T_0}, Y_{T_0})>\eps$, with probability at least $p_4\df
992: p_2p_3>0$, the processes $X$ and $Y$ will meet between times $T_0$
993: and the first time $R_1$ when they are both outside $\cT$. We
994: define for $k\geq 2$,
995: $$ S_k=S_1\circ \theta_{R_{k-1}}+R_{k-1} \qquad \hbox{and} \qquad
996: R_k=R_1\circ \theta_{S_k}+S_k.
997: $$
998: By the strong Markov property of $(X, Y)$,
999: $$ \P\left( \inf_{t\in [0, R_k)} \d(X_t, \, Y_t)\geq \eps \right) \leq (1-p_4)^k.
1000: $$
1001: Letting $k\to\infty$, we get
1002: $$ \P\left( \inf_{t\in [0, \infty)} \d(X_t, \, Y_t)
1003: \geq \eps \right) \leq 0
1004: $$
1005: for every $\eps>0$ and thus $(X,Y)$ is not a shy coupling.
1006: \qed
1007:
1008: \bigskip
1009: \noindent{\bf Example 3.7}. Suppose that $\cS$ is composed of a
1010: loop $\cU$ with a finite number of finite trees attached to it at
1011: points $x_k$, and the family $\{x_k\}$ is not rotation invariant
1012: in the following sense. We can assume without loss of generality
1013: that $\cU$ is isometric to the unit circle. For every $c\ne 0$,
1014: there exists $x_j$ such that $x_j e^{ic} \ne x_k$ for all $k$. See
1015: Fig.~3.5 for an example. We will show that in this case there is
1016: no shy coupling.
1017:
1018: \bigskip \vbox{ \epsfxsize=1.5in
1019: \centerline{\epsffile{shy8.eps}}
1020:
1021: \centerline{Figure 3.5.} }
1022: \bigskip
1023:
1024: Our argument will be very similar to that in Example 3.6. Recall
1025: the ``projection'' $Q$ from the previous example. We have $Q(x)=x$
1026: for $x\in \cU$ and $Q(x) = x_k$ if $x$ belongs to a tree that is
1027: attached to $\cU$ at $x_k$. Hence, $Q(X_t)$ may be regarded as a
1028: continuous process on the unit circle. We now choose a (random)
1029: continuous function $\Theta_X: [0,\infty) \to \R$ so that $Q(X_t)
1030: = e^{i\Theta_X(t)}$ for all $t\geq 0$, in the complex notation. We
1031: define $\Theta_Y$ in an analogous way. Note that $\Theta_X$ and
1032: $\Theta_Y$ are martingales. Therefore, $Z_t \df \Theta_X(t) -
1033: \Theta_Y(t)$ is also a martingale. We can now repeat the argument
1034: from Example 3.6 to show that there does not exist a shy coupling.
1035: \qed
1036:
1037: \bigskip
1038: \noindent{\bf Example 3.8}. Examples 3.3 and 3.5-3.7 may appear to
1039: suggest that if a graph has a vertex with degree 1 then a shy
1040: coupling exists only if there exists an isometry of $I:\cS\to \cS$
1041: with no fixed points. We will show that this is not the case. Our
1042: example is illustrated in Fig.~3.6. In this case, every isometry
1043: $I:\cS\to\cS$ has a fixed point. Nevertheless, we will show that
1044: there is a shy coupling in $\cS$.
1045:
1046: \bigskip \vbox{ \epsfxsize=3.0in
1047: \centerline{\epsffile{shy5.eps}}
1048:
1049: \centerline{Figure 3.6.} }
1050: \bigskip
1051:
1052: We will describe below the transition mechanism for $(X,Y)$ on
1053: some random intervals of time. We will assume that the transition
1054: probabilities of $(Y,X)$ are the same as those of $(X,Y)$. Hence,
1055: there is no need to describe cases symmetric to those discussed
1056: below, in the sense that the initial positions of $X$ and $Y$ are
1057: interchanged.
1058:
1059: Suppose that all edges $A_1, A_2, \dots, A_7$ have the same
1060: length, say 1. We assume that $X_0 = x_2$ and $Y_0 = x_3$.
1061:
1062: (i) Suppose that for some stopping time $T_1$ we have $X_{T_1} =
1063: x_2$ and $Y_{T_1} = x_3$. Then we let $T_2 = \inf\{t\geq T_1: X_t
1064: \notin (A_1 \cup A_2 \cup A_3)\setminus \{x_1\}\}$. Let $I: (A_1
1065: \cup A_2 \cup A_3)\setminus \{x_1,x_3\} \to (A_3 \cup A_4 \cup
1066: A_5)\setminus\{x_2,x_4,x_5\}$ be the one-to-one isometry
1067: satisfying $I(x_2) = x_3$, $I(A_1) = A_3$, $I(A_2) = A_4$ and
1068: $I(A_3) = A_5$. We let $Y_t = I(X_t)$ for $t\in[T_1,T_2]$. Note
1069: that at the stopping time $T_2$, we have one of the following
1070: configurations of the two particles: $(X_{T_2}, Y_{T_2})=(x_1,
1071: x_4)$, or $(X_{T_2}, Y_{T_2})= (x_1, x_2)$, or $(X_{T_2},
1072: Y_{T_2}) = (x_3, x_5)$.
1073:
1074: (ii) Suppose that for some stopping time $T_3$ we have $X_{T_3} =
1075: x_1$ and $Y_{T_3} = x_4$. Let $\{X_t, t\in [T_3, T_4]\}$ be
1076: Brownian motion on $\cS$ independent of the past with
1077: $X_{T_3}=x_3$, where $T_4 = \inf\{t\geq T_3: X_t = x_2\}$. For
1078: $t\in[T_3, T_4]$, we let $Y_t \in A_4$, with $\d(Y_t, x_4) =
1079: \d(X_t, x_1)$. Note that $X_{T_4} = x_2$ and $Y_{T_4} = x_3$.
1080:
1081: (iii) Suppose that for some stopping time $T_5$ we have $X_{T_5} =
1082: x_1$ and $Y_{T_5} = x_2$. Let $\{Y_t, t\in [T_5, T_6]\}$ be
1083: Brownian motion on $\cS$ independent of the past with
1084: $Y_{T_5}=x_2$,
1085: where $T_6 = \inf\{t\geq T_5: Y_t = x_1 \hbox{ or } x_3\}$. We
1086: label excursions of $Y$ from $x_2$ that stay in $A_3$ with marks
1087: ``1'' or ``2'', with equal probabilities, in such a way that the
1088: label of any excursion is independent of all other labels. Then we
1089: let $X_t$ be defined for $t\in [T_5, T_6]$ by $\d(X_t, x_1) =
1090: \d(Y_t, x_2)$ and the following conditions. If $Y_t \in A_1$ then
1091: $X_t \in A_2$, if $Y_t \in A_2$ then $X_t \in A_1$, if $Y_t \in
1092: A_3$ and $t$ belongs to an excursion marked ``1'' then $X_t \in
1093: A_1$, and if $Y_t \in A_3$ and $t$ belongs to an excursion marked
1094: ``2'' then $X_t \in A_2$. At time $T_6$ we have $(X_{T_6},
1095: Y_{T_6})= (x_2, x_1)$ or $(X_{T_6}, Y_{T_6}) =(x_2, x_3)$.
1096:
1097: Note that $\cS$ is symmetric with respect to the line containing
1098: $A_4$. If for some stopping time $T_7$ we have $( X_{T_7},
1099: Y_{T_7}) = (x_3, x_5)$, or $( X_{T_7}, Y_{T_7}) =(x_4, x_6)$, or
1100: $( X_{T_7}, Y_{T_7}) = (x_5, x_6)$, or one of these conditions is
1101: satisfied with the roles of $X$ and $Y$ interchanged, then we
1102: define the coupling on an appropriate random interval in a way
1103: analogous to that in (i)-(iii), using the symmetry of $\cS$.
1104:
1105: The above definitions for the ``local'' behavior of the coupling
1106: and the strong Markov property can now be used to define a process
1107: $(X_t,Y_t)$ for all $t\geq 0$. It is easy to see that the stopping
1108: times analogous to $T_1$, $T_3$ and $T_5$ will not have a finite
1109: point of accumulation. It is also easy to check that almost surely
1110: $ \d(X_t, Y_t) =1 $ for every $t>0$.
1111: \qed
1112:
1113: \bigskip
1114: \noindent {\bf 4. Reflected Brownian motion in Euclidean domains}.
1115:
1116: This section is the closest in spirit to the papers and problems
1117: which inspired the present research project. Suppose that
1118: $D\subset \R^d$ is a bounded connected open set which is either
1119: convex or has a $C^2$ boundary.
1120: We will consider couplings $(X,Y)$ of reflected Brownian motions
1121: in $D$, defined as follows. Let $\n(x)$ denote the unit inward
1122: normal vector at $x\in\prt D$. Let $B$ and $W$ be standard planar
1123: Brownian motions with $B_0=W_0=0$ defined on the same probability
1124: space and consider the following Skorohod equations,
1125: $$\eqalignno{
1126: X_t &= x_0 + B_t + \int_0^t \n(X_s) dL^X_s, &(4.1)\cr
1127: Y_t &= y_0 + W_t + \int_0^t \n(Y_s) dL^Y_s. &(4.2) }$$
1128: Here $L^X$ is the local time of $X$ on $\prt D$, i.e., a
1129: non-decreasing continuous process which does not increase when $X$
1130: is in $D$: $\int_0^\infty \bone_{D}(X_t) dL^X_t = 0$, a.s.
1131: Equation (4.1) has a unique pathwise solution $(X,L^X)$ such that
1132: $X_t \in \ol D$ for all $t\geq 0$ (see [Ta] when $D$ is convex
1133: domain and [LS] when $D$ is $C^2$ ). The ``reflected Brownian
1134: motion'' $X$ is a strong Markov process. We point out that $B$ is
1135: uniquely determined by $X$, and vice versa. The same remarks apply
1136: to (4.2), so, as a pair, $(X, Y)$ is also strong Markov.
1137:
1138: For a continuous semimartingale $M$, the symbol $\langle Z
1139: \rangle$ will stand for its quadratic variation process. When
1140: $M=(M^1, \cdots, M^d)$ and $N=(Z^1, \cdots , Z^d)$ are two
1141: continuous $\R^d$-valued semimartingales we will use $\langle M,
1142: N\rangle$ to denote $\sum_{i, j=1}^d \< M^i, N^j \>$. Note that
1143: the matrix-valued process $( \< M^i, N^j \>)_{1\leq i, j\leq d}$
1144: is non-negative definite and so $t\mapsto \< M , M\>_t$ is always
1145: non-decreasing. For $a, b \in \R^d$, we use $a\cdot b$ to denote
1146: the inner product between $a$ and $b$. We will use $\d(x, y)$ and
1147: $|x-y|$ interchangeably for the Euclidean distance between $x, y
1148: \in \R^d$.
1149:
1150: \bigskip
1151:
1152:
1153:
1154: \noindent{\bf Theorem 4.1}. {\sl Assume that $D\subset \R^d$ is a
1155: bounded convex domain. Let $X$ and $Y$ be two reflecting Brownian
1156: motion on $D$ given by (4.1)-(4.2).
1157:
1158: \item{(i)} Suppose that there is a strictly increasing function
1159: $\varphi$ with $\varphi (0)=0$ such that
1160: $$ d\< |X-Y|^2\>_t \geq \varphi (|X_t-Y_t| )\, d t \qquad \hbox{for }
1161: t<\sigma_0,
1162: $$
1163: where $\sigma_0:=\inf \{t>0: X_t=Y_t\}$. Then $(X,Y)$ is not a shy
1164: coupling.
1165:
1166: \item{(ii)} Suppose
1167: $D$ is strictly convex.
1168: Assume that $\< X-Y, X-Y\>_t$ (this is the same as $\< B-W,
1169: B-W\>_t)$) has a sublinear growth rate as $t\to \infty$, that is,
1170: $$\lim_{t\to \infty} \< X-Y, \, X-Y\>_t/t=0 \qquad \hbox{almost surely}.
1171: $$
1172: Then $(X,Y)$ is not a shy coupling.
1173:
1174: }
1175:
1176: \bigskip
1177: \noindent{\bf Proof}. Note that
1178: $$ X_t-Y_t=X_0-Y_0+(B-W)+\int_0^t \n (X_s) dL^X_s -\int_0^t \n (Y_s) dL^Y_s
1179: $$
1180: is a semimartingale. Define $R_t:=|X_t-Y_t|^2$. By Ito's formula,
1181: $$\eqalignno{
1182: d R_t &= 2 (X_t-Y_t) \cdot d(X_t-Y_t) + d\<X-Y, X-Y\>_t \cr
1183: &= 2(X_t-Y_t) \cdot d(B_t-W_t) - 2 (Y_t-X_t)\cdot \n (X_t) dL^X_t
1184: - 2(X_t-Y_t) \cdot \n (Y_t) dL^Y_t \cr & \hskip 0.2truein +
1185: d\<X-Y, X-Y\>_t. &(4.3) \cr}
1186: $$
1187:
1188: \medskip
1189:
1190: (i) Let
1191: $a>0$ be a constant whose value will be
1192: chosen in a moment and $f(r) = -r^{-a}$ for $r>0$. Then $f'(r) =
1193: ar^{-a-1}>0$ and $f''(r) = (-a-1)a r^{-a-2}<0$ for $r>0$.
1194: Define $U_t:= f(R_t)=f (|X_t-Y_t|^2)$. By Ito's formula, we have
1195: $$
1196: dU_t = f'(R_t) dR_t + {1\over 2} f'' (R_t) d\<R\>_t =dM_t+dV_t,
1197: $$
1198: where
1199: $$ dM_t = 2 aR_t^{-a-1} 2(X_t-Y_t) \cdot d(B_t-W_t)
1200: $$
1201: and
1202: $$\eqalignno{
1203: dV_t &= - 2 aR_t^{-a-1} \left( (Y_t-X_t) \cdot \n (X_t) dL^X_t
1204: + (X_t-Y_t)
1205: \cdot
1206: \n (Y_t) dL^Y_t \right) \cr & \hskip 0.2truein + a R_t^{-a-1}
1207: d\<X-Y, X-Y\>_t -2 a(a+1) R_t^{-a-2} d\< |X-Y|^2\>_t &(4.4) \cr}
1208: $$
1209: are the local martingale and bounded variation
1210: parts,
1211: respectively. We claim that for every $\eps >0$,
1212: $$ T_\eps := \inf \left\{ t>0: \ |X_t-Y_t | \leq \eps \right\}
1213: $$
1214: is finite almost sure.
1215: Suppose that $\P (T_\eps = \infty )>0$ for some $\eps >0$. We will
1216: show that this leads to a contradiction.
1217:
1218: Since $D$ is a convex domain, for $\sigma$-a.e. $x\in \partial D$,
1219: $\n(x)$ is well defined and
1220: $$ (y-x) \cdot \n (x) \geq 0 \qquad \hbox{for every }
1221: y
1222: \in \ol D.
1223: $$
1224: Note that the local time $L^X$ (respectively, $L^Y$) does not
1225: increase when $X$ (respectively $Y$) is on a subset of $\partial
1226: D$ having zero Lebesgue surface measure. Hence, (4.4) yields
1227: $$
1228: dV_t \leq a R_t^{-a-1} d\<X-Y, X-Y\>_t -2 a(a+1) R_t^{-a-2} d\<
1229: |X-Y|^2\>_t . \eqno(4.5)
1230: $$
1231: Note that $ d\<X-Y, X-Y\>_t = d \<B-W, B-W\>_t \leq 4dt$. This,
1232: (4.5) and the hypothesis in part (i) of this theorem imply that
1233: on $\{T_\eps = \infty\}$,
1234: $$\eqalignno{
1235: dV_t &\leq 4 a R_t^{-a-1}dt -2 a(a+1)
1236: R_t^{-a-2} \varphi (\eps) d t \cr
1237: &\leq -2a R^{-a-2} \left( (a+1) \varphi (\eps ) - 2 R_t \right) dt \cr
1238: & \leq -2a \eps^{-a-2}
1239: \left( (a+1) \varphi (\eps ) -2 \hbox{\rm diam} (D) \right)dt . \cr}
1240: $$
1241: For a fixed $\eps>0$, we can find $a>0$ sufficiently large so that
1242: for some $\lambda >0 $,
1243: $$ dV_t \leq - \lambda dt \qquad \hbox{for every } t>0
1244: \hbox{ on } \{ T_\eps = \infty\}.
1245: \eqno(4.6)
1246: $$
1247:
1248: The continuous local martingale $M$ is a time change of Brownian
1249: motion. By the law of iterated logarithm for Brownian sample path,
1250: for almost all $\omega \in \{T_\eps =\infty \}$, there is an
1251: unbounded increasing sequence $\{t_k, k\geq 1\}$ such that
1252: $\sup_{k\geq 1} | M_{t_k}(\omega) | <\infty $. This and (4.6)
1253: imply that $U_{t_k} (\omega)= V_0(\omega)+M_{t_k}(\omega)+
1254: V_{t_k}(\omega)$ tends to $-\infty$ as $k\to \infty$ on $\{T_\eps
1255: =\infty \}$ a.s. Consequently, $|X_{t_k}-Y_{t_k}|$ goes to $0$ as
1256: $k\to \infty$ on $ \{T_\eps =\infty \}$ a.s., which is a
1257: contradiction. This proves that particles $X$ and $Y$ come
1258: arbitrarily close to each other in finite time and, therefore, $X$
1259: and $Y$ is not a shy coupling.
1260:
1261: \medskip
1262:
1263: (ii) Now assume
1264: that $D$ is bounded and strictly convex
1265: and $\<X-Y, X-Y\>_t$ has a sublinear growth as $t\to \infty$.
1266: The strict convexity implies (in fact, it is equivalent to) the
1267: following condition. For every small $\eps >0$, there is a
1268: constant $a_\eps>0$ such that
1269: $$ (y-x) \cdot \n (x) \geq a_\eps \, |x-y| \qquad \hbox{for every }
1270: x\in \partial D \hbox{ and } y \in \ol D \hbox{ with } |x-y| \geq
1271: \eps. \eqno(4.7)
1272: $$
1273: Let $\sigma$ denote the surface measure on $\partial D$. Since
1274: reflecting Brownian motion in $D$ is a recurrent Feller process,
1275: it follows from the Ergodic Theorem that
1276: $$ \lim_{t\to \infty} {L^X\over t}= {\sigma (\partial D ) \over 2 |D|}
1277: = \lim_{t\to \infty} {L^Y \over t} \qquad \hbox{almost surely.}
1278: $$
1279: For every $\eps >0$, define $T_\eps:=\inf\{t>0: |X_t-Y_t| \leq \eps\}$.
1280: On $\{T_\eps =\infty\}$, we have from above
1281: and (4.7)
1282: that
1283: $$\eqalignno{
1284: & \liminf_{t\to \infty} {1\over t}
1285: \left(-2 \int_0^t (Y_t-X_t)\cdot \n(X_t) dL^X_t
1286: - 2 \int_0^t (X_t-Y)\cdot \n(Y_t) dL^Y_t + \<X-Y, X-Y\>_t \right) \cr
1287: & \ \ \ \leq \liminf_{t\to \infty} {1\over t} \left( -2 \eps
1288: a_\eps L^X_t - 2 \eps a_\eps L^Y_t + \<X-Y, X-Y\>_t \right) =
1289: -{ 2 \eps a_\eps \sigma ( \partial D) \over |D|}<0 . &(4.8)\cr}
1290: $$
1291: On the other hand, $M_t:=2\int_0^t (X_t-Y_t) \cdot d(B_t-W_t)$ is
1292: a continuous martingale and thus is a time-change of
1293: one-dimensional Brownian motion. By the law of iterated logarithm
1294: for Brownian sample path, for almost all $\omega \in \{T_\eps
1295: =\infty \}$, there is an unbounded increasing sequence $\{t_k,
1296: k\geq 1\}$ such that $\sup_{k\geq 1} | M_{t_k}(\omega) | <\infty
1297: $. This, (4.3) and (4.8) imply that $\lim_{t\to \infty}R_t=-\infty
1298: $ a.s. on $\{T_\eps =\infty\}$. Since $R_t \geq 0$, we conclude
1299: that $\P(T_\eps =\infty)=0$ for every $\eps>0$ and so $(X, Y)$ is
1300: not a shy coupling.
1301: \qed
1302:
1303: \bigskip
1304: In the remainder of this section, we take $d=2$, but this is only
1305: for notational convenience. We will show in the next example that
1306: the method of proof of Theorem 4.1, based on the It\^o formula,
1307: does not extend to arbitrary couplings. The example may have some
1308: interest of its own. We will show in Theorem 4.3 below that, in
1309: fact, there is no shy coupling of reflecting Brownian motions on
1310: any bounded $C^1$-smooth strictly convex domain.
1311:
1312: \bigskip
1313:
1314: \noindent{\bf Example 4.2}. We will show that there exist planar
1315: Brownian motions $B$ and $W$ with the property that $\d(B_t,W_t) =
1316: \sqrt{2t + \d(B_0, W_0)^2}$ for $t\geq 0$, assuming that $B_0\ne
1317: W_0$. In particular, the distance between the two processes grows
1318: in a deterministic way.
1319:
1320: Suppose that $(B, W)$ has the above mentioned property with $B_0$
1321: and $W_0$ taking values in $\ol D$. Let $X$ and $Y$ be the
1322: pathwise solutions
1323: of (4.1)-(4.2) but with the above $B$ and $W$ in place of $x_0+B$
1324: and $y_0+W$ there.
1325: We have
1326: $$ d\< |X_t-Y_t|^2\>= d\< |B_t-W_t|^2\>=0,
1327: $$
1328: while
1329: $$ d\<X-Y, X-Y\>_t=d\<B-W, B-W\>_t =d \left( |B_t-W_t|^2\right)=2t.
1330: $$
1331: So $V_t$ in (4.4) becomes
1332: $$\eqalignno{
1333: dV_t &= - 2 aR_t^{-a-1} \left( (Y_t-X_t) \cdot \n (X_t) dL^X_t
1334: + (X_t-Y_t)
1335: \cdot
1336: \n (Y_t) dL^Y_t \right) \cr & \hskip 0.2truein + 2 a R_t^{-a-1} d
1337: t. \cr}
1338: $$
1339: Hence the method used in the proof of Theorem 4.1 does not work for
1340: this coupling
1341: $(X , Y$). Moreover, since $|X_t-Y_t|$ grows deterministically
1342: when both $X_t$ and $Y_t$ are away from the
1343: boundary and decreases when one of them is on the boundary,
1344: neither $|X_t-Y_t|$ nor any deterministic monotone
1345: function of $|X_t-Y_t |$ is a submartingale or a supermartingale.
1346:
1347:
1348: We now present the construction of $B$ and $W$ with the properties
1349: mentioned above.
1350: Let $B$ be a Brownian motion in $\R^2$ starting from $x_0$. For a
1351: vector $v=(a, b)\in \R^2$, we use $v^\perp$ to denote its
1352: orthogonal vector $(b, -a)$. Let $y_0 \in \R^2$ be a point
1353: different from $x_0$. Consider the following SDE for $W$ in $\R^2$
1354: with $W_0=y_0$:
1355: $$ dW_t= {1\over |W_t-B_t|^2} \left( \left( (W_t-B_t)\cdot dB_t \right) (W_t-B_t)
1356: - \left( (W_t-B_t)^\perp \cdot dB_t \right) (W_t-B_t)^\perp \right) .
1357: $$
1358: In words, at any given time $t>0$, $W_t$ takes
1359: a synchronous step
1360: with $B_t$ along the direction $W_t-B_t$, while $W_t$ moves in the
1361: opposite direction but with the same magnitude as $B_t$ along the
1362: perpendicular direction $(W_t-B_t)^\perp$. The above SDE for $W$
1363: has a unique solution up to $\tau:=\inf\{t>0: W_t=B_t\}$, since
1364: the diffusion coefficients are $C^\infty$ up to that time. It can
1365: be computed directly that $d\left( |W_t-B_t|^2 \right)= 2 dt$ and
1366: consequently $|W_t-B_t|^2= |x_0-y_0|^2+2t$. So $\tau =\infty$.
1367: It is standard to check
1368: that $W=(W^1, W^2)$ is a continuous local martingale with $\< W^i,
1369: W^i\>_t=t$ for $i=1, 2$ and $\<W^1, W^2\>=0$. Therefore $W$ is a
1370: Brownian motion in $\R^2$ starting from $y_0$. \qed
1371:
1372:
1373:
1374: \bigskip
1375:
1376: We will show next that
1377: in a $C^1$-smooth strictly convex domain $D$, every coupling of
1378: reflecting Brownian motions on $\ol D$ must come arbitrarily close
1379: to each other
1380: in finite time.
1381:
1382: \bigskip
1383:
1384: \noindent{\bf Theorem 4.3}. {\sl Suppose that $D$ is a bounded
1385: convex planar domain with a
1386: $C^1$-smooth
1387: boundary that does not
1388: contain any line segments. Then there does not exist a shy
1389: coupling $(X,Y)$ of reflected Brownian motions in $D$.
1390:
1391: }
1392:
1393: \bigskip
1394: \noindent{\bf Proof}. The idea of the proof is inspired by
1395: differential games of pursuit (see [F]). We will show that with
1396: positive probability, one of the particles will pursue the other
1397: one in such a way that the distance between the two particles
1398: decreases either because the diffusion component of the second
1399: process does not move the second particle sufficiently fast or the
1400: second particle hits the boundary and is pushed back towards the
1401: first one.
1402:
1403:
1404:
1405: \medskip
1406:
1407: {\it Step 1}. We will define several constants $\eps_k$ in this
1408: step. The definitions will be labeled (a), (b), (c), etc. Each of
1409: these definitions is really a simple lemma asserting the existence
1410: of a constant with stated properties. Since the proofs do not need
1411: more than high school geometry, we omit most of the proofs. The
1412: constants $\eps_k$ are defined relative to each other, but
1413: $\eps_k$ may depend only on the values of $\eps_j$ for $j<k$.
1414:
1415: (a) Let $\eps_0>0$ be so small that for every $x\in D$ with
1416: $\d(x,\prt D) \leq \eps_0$ there exists a unique point in $\prt D$
1417: whose distance from $x$ is minimal.
1418:
1419: We make $\eps_0>0$ smaller, if necessary, so that the following is
1420: true. Consider any point $y\in \prt D$ and let $CS_1$ be the
1421: orthonormal coordinate system such that $y=0 \in \prt D$ and
1422: $\n(0) $ lies on the second axis. Write $\n(x) = (\n_1(x),
1423: \n_2(x))$. Then $|\n_1(x)| \leq \n_2(x)/100$ for $x \in \prt D
1424: \cap \cB(0, \eps_0)$ in $CS_1$.
1425:
1426: We fix an arbitrary $\eps_1 \in (0, \eps_0]$. It will suffice to
1427: prove that for any $x_0,y_0 \in \ol D$, if $(X_0,Y_0)=(x_0,y_0)$
1428: then, with probability one, there exists $t< \infty$ such that
1429: $\d(X_t,Y_t) \leq \eps_1$.
1430:
1431: (b) The angle between two vectors will be denoted $\angle
1432: (\cdot\,,\,\cdot)$, with the convention that it takes values in
1433: $(-\pi, \pi]$. Since $D$ is a bounded and strictly convex domain,
1434: there exists $\eps_2 \in (0, \pi /2)$ such that for every $x\in
1435: \prt D$ and $y\in \ol D$ satisfying $\d(x,y) \geq \eps_1/2$,
1436: $$ \angle (\n(x), y-x ) \in [ - \pi/2 +\eps_2, \, {\pi / 2}-\eps_2 ].
1437: \eqno(4.9)
1438: $$
1439:
1440: (c) Let $\cL(x,r) $ be the cone spanned by $ \{\n (y)$, $y\in \prt
1441: D\cap \cB(x, r)\}$. Since $\prt D$ is $C^1$-smooth, $\cL(x,r)$ is
1442: a wedge. Hence, all linear combinations of vectors in $\cL(x,r)$
1443: with non-negative coefficients belong to $\cL(x,r)$. An easy
1444: approximation argument shows that if $X_t \in \cB(x,r)$ for all
1445: $t\in(s,u)$ then $\int_s^u \n(X_t) dL^X_t \in \cL(x,r)$.
1446:
1447: We will now choose $\eps_3\in(0,\eps_1/8)$. Consider vectors $\v$
1448: and $\w$ satisfying the following conditions, relative to $x_0,
1449: y_0$, and $\eps_3$.
1450:
1451: If $\d(x_0,\prt D)\leq \eps_3$ then $ \v \in \cL(x_0, 2\eps_3)$
1452: and $|\v| \leq \eps_3$. If $\d(x_0,\prt D)> \eps_3$ then $ \v =
1453: 0$.
1454:
1455: If $\d(y_0,\prt D)\leq \eps_3$ then $ \w \in \cL(y_0, 2\eps_3)$
1456: and $|\w| \leq \eps_3$. If $\d(y_0,\prt D)> \eps_3$ then $ \w =
1457: 0$.
1458:
1459: We will show that (4.9) implies that we can find sufficiently
1460: small $\eps_3>0$ so that the following is true. Suppose that $x_0,
1461: y_0 \in \ol D$ with $\d(x_0, y_0) \geq \eps_1$. Assume that
1462: $x_1\in \cB(x_0, 2\eps_3)$ and $y_1\in \cB(y_0, 2\eps_3)$. Then
1463: $$\d(x_1 + \v, y_1 +\w) \leq \d(x_1, y_1).\eqno(4.10)$$
1464: To see this, choose some $x_2$ and $y_2$ so that the following
1465: conditions hold.
1466:
1467: If $\d(x_0,\prt D)\leq \eps_3$ then $x_2\in \prt D \cap \cB(x_0, 2\eps_3)$
1468: with $ \v =c \n(x_2)$.
1469: If $\d(x_0,\prt D)> \eps_3$ then $ x_2 = x_ 0$.
1470:
1471: If $\d(y_0,\prt D)\leq \eps_3$ then $y_2\in \prt D \cap \cB(y_0, 2\eps_3)$
1472: with $ \w =c \n(y_2)$.
1473: If $\d(y_0,\prt D)> \eps_3$ then $ y_2 = y_ 0$.
1474:
1475: Since $\eps_3< \eps_1 /8$,
1476: $$ |x_2-y_2|\geq |x_0-y_0|-|x_0-x_2|-|y_0-y_2| \geq \eps_1 -4 \eps_3
1477: \geq \eps_1 /2.
1478: $$
1479: Thus by (4.9), we have
1480: for sufficiently small $\eps_3>0$,
1481: $$ \eqalignno{
1482: &\d(x_1 + \v, y_1 +\w)^2 \cr &\leq |x_1-y_1|^2+|\v-\w|^2+2
1483: (x_1-y_1)\cdot (\v-\w) \cr &\leq |x_1-y_1|^2+|\v-\w|^2+2
1484: (x_2-y_2)\cdot (\v-\w) + 2(|x_1-x_2|+|y_1-y_2|)|\v-\w|\cr &\leq
1485: |x_1-y_1|^2+ |\v -\w|^2 -2 (\sin \eps_2) |x_2-y_2| ( |\v|+|\w|) +
1486: 4 (8\eps_3)|\v-\w| \cr
1487: & \leq |x_1-y_1|^2 + (|\v|+|\w|) \left( 6\eps_3 -2 (\sin \eps_2)
1488: (\eps_1/2)
1489: + 32 \eps_3 \right) \cr
1490: & =
1491: |x_1-y_1|^2 - (|\v|+|\w|) \left(
1492: \eps_1 \sin \eps_2
1493: -38 \eps_3 \right) &(4.11)\cr & \leq \d(x_1,
1494: y_1)^2 . \cr}
1495: $$
1496:
1497: (d) Since $D$ is bounded, we can find $\eps_4>0$ and $N<\infty$,
1498: such that if $x_1, x_2, \dots$ is a sequence of points with
1499: $x_1\in \ol D$, $\d(x_k,x_{k-1})\geq \eps_3/8$ and $|\angle(x_k -
1500: x_{k-1}, x_{k+1} - x_k)| \leq \eps_4$ for all $k$ then $x_N\notin
1501: \ol D$.
1502:
1503: (e) We choose $\eps_5, \eps_6 >0$ so that the following is true.
1504: Suppose that $x_0,y_0,x_1,y_1$ and $x_2$ satisfy the conditions
1505: $\d(x_0,y_0)\geq \eps_1$, $\d(x_1,y_1)\geq \eps_1$ and
1506: $$\eqalignno{
1507: &|\angle (x_1- y_1, x_0-y_0)| \leq \eps_5,\cr
1508: &\d(x_1 ,
1509: x_0 + (\eps_3/4) (y_0-x_0)/\d(x_0,y_0) ) \leq \eps_6,\cr
1510: &\d(x_2 ,
1511: x_1 + (\eps_3/4) (y_1-x_1)/\d(x_1,y_1) ) \leq \eps_6.\cr}$$
1512: Then $|\angle(x_1 - x_{0}, x_{2} - x_1)| \leq \eps_4$.
1513:
1514: We make $\eps_6 $ smaller, if necessary, so that $\eps_6 <
1515: \eps_3/8$.
1516:
1517: (f) We make $\eps_6>0$ smaller, if necessary so that the following
1518: holds. Suppose that $x_0,y_0,x_1$ and $y_1$ satisfy the conditions
1519: $\d(x_0,y_0)\geq \eps_1$,
1520: $$\eqalignno{
1521: &\d(x_1 , x_0 + (\eps_3/4) (y_0-x_0)/\d(x_0,y_0) ) \leq
1522: \eps_6,&(4.12)\cr
1523: &\d(y_1 , y_0 + (\eps_3/4) (y_0-x_0)/\d(x_0,y_0) ) \leq \eps_6.&(4.13)\cr
1524: }$$
1525: Then $|\angle(x_1 - y_1, x_0 - y_0)| \leq \eps_5$.
1526:
1527: (g) We can find $\eps_7,\eps_8 >0$ with the following properties.
1528: Suppose that $x_0,y_0\in \ol D$, $x_1,x_2\in \R^2$,
1529: $\d(x_0,y_0)\geq \eps_1$ and the following conditions are
1530: satisfied.
1531:
1532: If $\d(x_0,\prt D)\leq \eps_3$ then $ \v \in \cL(x_0, 2\eps_3)$
1533: and $|\v| \leq \eps_3$. If $\d(x_0,\prt D)> \eps_3$ then $ \v =
1534: 0$.
1535:
1536: If $\d(y_0,\prt D)\leq \eps_3$ then $ \w \in \cL(y_0, 2\eps_3)$
1537: and $|\w| \leq \eps_3$. If $\d(y_0,\prt D)> \eps_3$ then $ \w =
1538: 0$.
1539:
1540: Assume that
1541: $$\eqalignno{
1542: &\d(x_1 , x_0 + (\eps_3/4) (y_0-x_0)/\d(x_0,y_0) ) \leq
1543: \eps_8,&(4.14)\cr
1544: &\d(y_1 , y_0 + (\eps_3/4) (y_0-x_0)/\d(x_0,y_0) ) \geq \eps_6/2,&(4.15)\cr
1545: & \d(y_1, y_0) \leq \eps_3/4 +\eps_8. &(4.16) \cr
1546: }$$
1547: Then $\d(x_1+\v, y_1+\w) \leq \d(x_0,y_0) -\eps_7$.
1548:
1549: (h) It is easy to see from (4.11) that we can strengthen (4.10) as
1550: follows. We can make $\eps_7>0$ smaller, if necessary, so that if
1551: $|\v| \geq \eps_6/2$ or $|\w| \geq \eps_6/2$, and the assumptions
1552: stated in Step 1(c) hold then
1553: $$\d(x_1 + \v, y_1 +\w) \leq \d(x_1, y_1)-2\eps_7.\eqno(4.17)$$
1554:
1555: \medskip
1556:
1557: \noindent{\it Step 2}. Suppose that $X_0, Y_0 \in \ol D$ with
1558: $\d(X_0,Y_0) \geq \eps_1$. Consider the following events,
1559: $$\eqalign{
1560: F_1(t)&=\{
1561: \d(X_{t}, Y_{t}) \leq \d(X_0,Y_0) - \eps_7\},\cr
1562: F_2(t) & =
1563: \{|\angle (X_{t}- Y_{t}, X_0-Y_0)| \leq \eps_5\} ,\cr
1564: F_3(t) &=
1565: \{\d(X_t, X_0 + (\eps_3/4)
1566: (Y_0-X_0)/\d(X_0,Y_0) ) \leq \eps_6\land \eps_8\}, \cr
1567: F_4(t) & =\{ \d(X_{t}, Y_{t}) \leq \d(X_0,Y_0) +
1568: \eps_7/(4N)\},\cr
1569: F_5(t) & = (F_1(t) \cup F_2(t) ) \cap F_3(t) \cap F_4(t).
1570: }$$
1571: We will show in Step 4 that $\P( F_5(t_1)) > p_1$ for some $t_1,
1572: p_1>0$ that do not depend on $X_0$ and $Y_0$.
1573:
1574: Let $\eps_9 = \eps_7/(16N) \land \eps_3/8\land \eps_6/8\land
1575: \eps_8/5$. Recall that $B$ and $W$ are Brownian motions with
1576: $B_0=W_0=0$ driving $X$ and $Y$ in the sense of (4.1)-(4.2) and
1577: let
1578: $$\eqalign{
1579: A_1(t)&=\left\{B_t\in \cB \left( (\eps_3/4) (Y_0-X_0)/\d(X_0,Y_0), \
1580: \eps_9 \right) \right\},\cr
1581: A_2(t)&=\left\{\sup_{ s\in[0,t]}
1582: | B_s - (s/t) B_t | \leq \eps_9 \right\},\cr
1583: A_3(t)&= \left\{ |W_t| \leq \eps_3/4 + \eps_9 \right\},\cr
1584: A_4(t) & = \left\{\sup_{ s\in[0,t]} | W_s - (s/t) W_t| \leq \eps_9
1585: \right\},\cr
1586: A_5(t) &= A_1(t) \cap A_2(t) \cap A_3(t) \cap A_4(t).
1587: }$$
1588: We will argue in the rest of this step of the proof that
1589: $\P(A_5(t_1) )> p_1$ for some $p_1,t_1>0$. In later steps, we will
1590: show that $A_5(t)\subset F_5(t)$.
1591:
1592: Recall that $B$ is a two-dimensional Brownian motion with $B_0=0$
1593: and let $T_r = \inf\{t\geq 0: |B_t| > r\}$. Note that, by Brownian
1594: scaling,
1595: $$ \P(T_r < t)
1596: = \P\left(\max_{0\leq s \leq t}|B_s| >r \right)
1597: = \P\left( \max_{0\leq s \leq 1}|B_s| > r/\sqrt{t} \right).
1598: $$
1599: By the large deviations principle (see [RY], Ch. VIII, Thm. 2.11),
1600: $$\lim_{t/r^2 \to 0} (2t/r^2) \log \P(T_r < t\mid B_0=0)
1601: = -1.\eqno(4.18)$$
1602:
1603: Let
1604: $$\eqalign{
1605: r_0 &= \eps_3/4, \cr
1606: A_6(t) &= \{ T_{r_0} < t\}, \cr
1607: A_7 & = \{B_{T_{r_0}} \in \cB ( (\eps_3/4) (Y_0-X_0)/\d(X_0,Y_0),
1608: \ \eps_9 /2)\}, \cr
1609: A_8(t) & = \left\{ \sup_{T_{r_0} \leq s \leq
1610: T_{r_0}+t} | B_s - B_{T_{r_0}} | \leq \eps_9/2 \right\}.
1611: }$$
1612: Clearly $A_6(t)$, $A_7$ and $A_8(t)$ are independent, and $ A_6(t)
1613: \cap A_7 \cap A_8(t)\subset A_1(t)$. So
1614: $$
1615: \P(A_1(t)) \geq \P_6(A_6(t)) \cdot \P(A_7) \cdot \P(A_8(t)).
1616: $$
1617: Note that
1618: $\lim_{t\to 0} P(A_8(t)) = 1$ by the strong Markov
1619: property applied at $T_{r_0}$ and the event $A_7$ is independent of $t$.
1620: This together with (4.18) yields
1621: $$\liminf_{t \to 0}\, t \log \P(A_1(t)) \geq -r_0^2/2= -(\eps_3/4)^2/2.$$
1622: We obtain directly from (4.18) that
1623: $$\limsup_{t \to 0} \, t \log \P(A^c_3(t)) \leq -(\eps_3/4+\eps_9)^2/2.$$
1624: This implies that for all sufficiently small $t>0$ we have
1625: $$\P(A_1(t) \cap A_3(t)) >0.\eqno(4.19)$$
1626:
1627: The process $\{B_s - (s/t)B_t, \, s\in[0,t]\}$ is Brownian bridge
1628: with duration $t$ seconds, i.e., Brownian motion starting from 0
1629: and conditioned to be at 0 at time $t$. If $\wt T_r$ denotes the
1630: hitting time of $r$ by the absolute value of the Brownian bridge
1631: then we have a formula analogous to (4.18), $\limsup_{t/r^2 \to 0}
1632: (2t/r^2) \log \P(\wt T_r < t) \leq -1$. Thus
1633: $$ \limsup_{t \to 0} t \log \P(A^c_2(t)) \leq -\eps_9^2/2.
1634: \eqno(4.20)
1635: $$
1636: Note that Brownian motion $B=\{B_t, t\geq 0\}$ is a Gaussian
1637: process. Since for every $0\leq s<t$, $B_s-(s/ t) B_t$ and $B_t$
1638: have zero covariance, the process $\{B_s - (s/t)B_t, \,
1639: s\in[0,t]\}$ is independent of $B_t$. Similar remarks apply to
1640: $W$. Hence
1641: $$
1642: \P( A_1(t) \cap A_2(t))= \P(A_1(t)) \P(A_2(t))
1643: \quad \hbox{ and } \quad
1644: \P( A_3(t) \cap A_4(t))= \P(A_3(t)) \P(A_4(t)).
1645: $$
1646: This, (4.19) and (4.20)
1647: imply that
1648: $$ \P(A_5(t_1) ) > p_1 \qquad \hbox{for some } t_1,p_1>0.
1649: \eqno (4.21)
1650: $$
1651: \medskip
1652:
1653: \noindent{\it Step 3}. Let $R_t = \int_0^t \n(X_s) dL^X_s$. We
1654: will show that if $A_1(t_0)\cap A_2(t_0)$ holds for some $t_0>0$
1655: then $|R_s| \leq 4\eps_9$ for every $s\in[0,t_0]$.
1656:
1657: Since $A_1(t_0)$ and $ A_2(t_0)$ hold, $|B_t| \leq \eps_3/4
1658: +2\eps_9 \leq \eps_3/2$ for every $t\leq t_0$. Thus when $\d(X_0,
1659: \prt D) \geq \eps_3$, $X_0+B_s \in \ol D$ for all $s\in [0,t_0]$.
1660: By the uniqueness of the solution to (4.1), $X_s =X_0+B_s$ for
1661: $s\in [0,t_0]$, and $R_s = 0$ for all $s\in[0,t_0]$.
1662:
1663: Suppose that $\d(X_0, \prt D) \leq \eps_3$. By the assumptions
1664: made in Step 1(a), there exists a unique point $y\in \prt D$ with
1665: the smallest distance to $X_0$. Let $CS_1$ be the orthonormal
1666: coordinate system such that $y=0 \in \prt D$ and $\n(0) $ lies on
1667: the second axis. Recall from Step 1(a) that $\n(x) = (\n_1(x),
1668: \n_2(x))$ and $|\n_1(x)| \leq \n_2(x)/100$ for $x \in \prt D \cap
1669: \cB(0, 3\eps_3)$ in $CS_1$. Write $R_t = (R^1_t, R^2_t)$. By the
1670: opening remarks in Step 1(c), $R_t \in \cL(0,r)$ if $X_s\in
1671: \cB(0,r)$ for all $s\leq t$. This implies that $|R^1_t| \leq
1672: R^2_t/100$ if $X_s\in \cB(0,3\eps_3)$ for all $s\leq t$.
1673:
1674: Let $T_1 = \inf\{t\geq 0: |R_{t}| > 4 \eps_9\}$. We will assume
1675: that $T_1 < t_0$ and show that this leads to a contradiction.
1676: Since $|B_t| \leq \eps_3/2$ for every $t\leq t_0$, we have $|B_t +
1677: R_t| \leq \eps_3/2 + 4\eps_9 \leq \eps_3$ for every $t\leq T_1$.
1678: Hence,
1679: $$
1680: |X_t| \leq | X_0 | + |B_t + R_t| \leq \eps_3 + \eps_3 =
1681: 2 \eps_3 \quad \hbox{ for every } t\leq T_1.
1682: $$
1683: It follows that $R^2_t\geq 0$ for every $t\leq T_1$ and
1684: $|R^1_{T_1}| \leq R^2_{T_1}/100$. Note that, by Step 1(a), the
1685: slope of the tangent line at points in $\cB(0, 3\eps_3)\cap
1686: \partial D$ is between $-1/100$ and $1/100$ and that $\d(X_{T_1} ,
1687: X_0 + B_{T_1})=|R_{T_1}|=4\eps_9$. The last observation and the
1688: fact that $X_{T_1} \in \prt D$ imply that $\d(X_0 + B_{T_1}, \prt
1689: D)\geq (2/3) |R_{T_1}|=8\eps_9/3$. Since $A_1(t_0)$ and $
1690: A_2(t_0)$ hold, $\d(X_0 +B_t, D) \leq 2\eps_9 $ for every $t\leq
1691: t_0$ and in particular for $t=T_1$. This contradiction proves the
1692: claim that $|R_s| \leq 4\eps_9$ for every $s\in[0,t_0]$.
1693:
1694: \medskip
1695:
1696: Let $\wt R_t = \int_0^t \n(Y_s) dL^Y_s$. We claim that if
1697: $A_3(t_0)\cap A_4(t_0)$ holds for some $t_0>0$ then $|\wt R_s|
1698: \leq \eps_3$ for every $s\in[0,t_0]$. To see this, observe that
1699: $\d(Y_0 +W_t, D) \leq \eps_3/4 + 2\eps_9 \leq \eps_3/2$ for $t\leq
1700: t_0$ and use that same argument as in the case of $R_t$.
1701:
1702: \medskip
1703:
1704: \noindent{\it Step 4}. Fix $t_0>0$. We will show that
1705: $A_5(t_0)\subset F_5(t_0)$. Assume that $A_5(t_0)$ holds.
1706:
1707: Since $A_1(t_0)$ holds, we have in view of Step 3,
1708: $$\eqalign{
1709: &\d(X_{t_0}, X_0 + (\eps_3/4)
1710: (Y_0-X_0)/\d(X_0,Y_0) ) \cr
1711: &\leq \d(B_{t_0}, (\eps_3/4)
1712: (Y_0-X_0)/\d(X_0,Y_0) ) + |R_{t_0}|
1713: \leq \eps_9 + 4\eps_9
1714: \leq \eps_6\land \eps_8.}$$
1715: In other words, $F_3(t_0)$ holds.
1716:
1717: Since $A_1(t_0)$ and $A_3(t_0)$ hold we have, using simple
1718: geometry,
1719: $$\d(X_0 + B_{t_0}, Y_0 + W_{t_0}) \leq \d(X_0,Y_0) + 3 \eps_9.\eqno(4.22)$$
1720: We will apply (4.10) with $x_1 = X_0 + B_{t_0}$, $y_1 = Y_0 +
1721: W_{t_0}$, $\v = R_{t_0}$ and $\w = \wt R_{t_0}$. We have assumed
1722: that $A_5(t_0)$ holds so $|B_{t}| \leq \eps_3$ and $|W_{t}| \leq
1723: \eps_3$ for $t\leq t_0$. This and Step 3 imply that for $t\leq
1724: t_0$,
1725: $$\d(X_t, X_0) \leq |B_t| + |R_t| \leq 2 \eps_3,$$
1726: and similarly $\d(Y_t, Y_0) \leq 2 \eps_3$. By the opening
1727: remarks in Step 1(c),
1728: $$ \v = \int_0^{t_0} \n(X_t) dL^X_t \in \cL(X_0, 2 \eps_3)
1729: \quad \hbox{ and } \quad \w = \int_0^{t_0} \n(Y_t) dL^Y_t
1730: \in \cL(Y_0, 2 \eps_3).
1731: $$
1732: By Step 3, $|\v| \leq \eps_3$ and $|\w| \leq
1733: \eps_3$. We have shown that all the conditions listed in Step 1(c)
1734: are satisfied so we can apply (4.10) to obtain
1735: $$\d(X_{t_0} , Y_{t_0}) \leq \d(X_0,Y_0) + 3 \eps_9.$$
1736: This proves that $F_4(t_0)$ holds.
1737:
1738: It will now suffice to show that if $F_2(t_0)$ does not hold then
1739: $F_1(t_0)$ does. Assume that $F_2(t_0)$ does not hold.
1740:
1741: If all of the following conditions hold,
1742: $$\eqalignno{
1743: & | B_{t_0} - (\eps_3/4)
1744: (Y_0-X_0 )/\d(X_0,Y_0)| \leq \eps_6/2\land c_8,&(4.23)\cr
1745: & | W_{t_0} - (\eps_3/4) (Y_0-X_0)/\d(X_0,Y_0) |
1746: \leq \eps_6/2,&(4.24)\cr
1747: &\d(X_{t_0}, X_0 + B_{t_0} ) \leq \eps_6/2,&(4.25)\cr
1748: &\d(Y_{t_0}, Y_0 + W_{t_0} ) \leq \eps_6/2,&(4.26)
1749: }$$
1750: then (4.12) and (4.13) hold with $(x_1, y_1) = (X_{t_0}, Y_{t_0})$
1751: and $(x_0, y_0)=(X_0, Y_0)$, and this implies $F_2(t_0)$, which is
1752: a contradiction. Hence, at least one of the conditions
1753: (4.23)-(4.26) must fail. The first of these conditions holds
1754: because $A_1(t_0)$ is true. By Step 3, (4.25) holds.
1755:
1756: Suppose that (4.26) fails. In view of (4.22), we can apply (4.17)
1757: to $x_1 = X_0 + B_{t_0}$, $y_1 = Y_0 + W_{t_0}$, $\v = R_{t_0}$
1758: and $\w = \wt R_{t_0}$ to obtain
1759: $$\d(X_{t_0} , Y_{t_0}) \leq \d(X_0,Y_0) + 3 \eps_9 -2\eps_7
1760: \leq \d(X_0,Y_0) - \eps_7.$$
1761: Hence, we have $F_1(t_0)$ in this case.
1762:
1763: Suppose that (4.24) fails. Then, in view of Step 3, (4.14)-(4.16)
1764: hold with $(x_0, y_0)=(X_0, Y_0)$, $(x_1, y_1)=(X_0 + B_{t_0}, Y_0
1765: + W_{t_0})$, $\v = R_{t_0}$ and $\w = \wt R_{t_0}$ and we have
1766: $$
1767: \d(X_{t_0} , Y_{t_0}) \leq \d(X_0,Y_0) - \eps_7.
1768: $$
1769: Hence, $F_1(t_0)$ holds. This proves that $A_5(t_0)\subset F_5
1770: (t_0)$.
1771:
1772: \medskip
1773:
1774: \noindent{\it Step 5}. Fix some $\eps_1\in (0,\eps_0)$ and let
1775: $\eps_j$'s be defined relative to $\eps_1$ as in Step 1. Let
1776: $\rho$ be the diameter of $D$ and let $N_0$ be an integer greater
1777: than $4\rho/\eps_7$. Recall from (4.21) in Step 2 that for some
1778: $p_1,t_1>0$ we have $\P(A_5(t_1) )> p_1$. Let
1779: $$S_1 = \inf\{t\geq 0: \d(X_t,Y_t) \leq \eps_1
1780: \hbox{ or } F_5(t) \hbox{ holds}\}\land (2t_1).
1781: $$
1782: By (4.21) and Step 4, $\P(S_1 \leq t_1) > p_1$. Recall that
1783: $\theta$ stands for the usual Markov shift and define
1784: $$ S_0=0 \qquad \hbox{and} \qquad
1785: S_k = S_1 \circ \theta_{S_{k-1}} + S_{k-1}\quad \hbox{for } k\geq 1.
1786: $$
1787: Recall integer $N$ defined in Step 1. By the strong Markov
1788: property, with probability no less than $p_1^{2NN_0}>0$, we have
1789: $S_k-S_{k-1} \leq t_1$ for all $k \leq 2NN_0$.
1790:
1791: We will argue that if $\bigcap_{k\leq 2NN_0} \{ S_k-S_{k-1} \leq
1792: t_1\}$ holds then $\d(X_t, Y_t) \leq \eps_1$ for some $t \leq
1793: 2NN_0 t_1$. Assume otherwise. Then $F_5(S_1) \circ \theta
1794: _{S_{k-1}}$ holds for every $k \leq 2NN_0$. In particular,
1795: $F_4(S_1) \circ \theta _{S_{k-1}}$ holds for every $k \leq 2NN_0$.
1796: Let $F_6(t) = F_2(t)\cap F_3(t)$. Since $F_5(t) \subset F_4(t)
1797: \cap ( F_1(t) \cup (F_2(t)\cap F_3(t)))$, for every $k \leq
1798: 2NN_0$, at least one of the events $F_1(S_1) \circ \theta
1799: _{S_{k-1}}$ and $F_6(S_1) \circ \theta _{S_{k-1}}$ holds.
1800:
1801: Consider any $j\leq N_0$. If $F_6(S_1) \circ \theta _{S_{k-1}}$
1802: holds for $k= 2jN, 2jN+1, \dots, 2(j+1)N -1$ then $X_{S_k}$'s and
1803: $Y_{S_k}$'s satisfy the following conditions for $k= 2jN, 2jN+1,
1804: \dots, 2(j+1)N -1$,
1805: $$\eqalign{
1806: \d\left (X_{S_{k+1}}, X_{S_k}+ (\eps_3/4)(Y_{S_k} - X_{S_k})/
1807: \d (X_{S_k},Y_{S_k}) \right) &\leq \eps_6, \cr
1808: \left|\angle \left( X_{S_k} - Y_{S_k}, X_{S_{k+1}} - Y_{S_{k+1}}
1809: \right) \right| &\leq \eps_5.
1810: }$$
1811: This implies, by Step 1(e), that for $k= 2jN, 2jN+1, \dots,
1812: 2(j+1)N -2$,
1813: $$
1814: |\angle(X_{S_{k+1}} - X_{S_k}, X_{S_{k+2}} - X_{S_{k+1}})|
1815: \leq \eps_4.
1816: $$
1817: Since $F_3(S_1) \circ \theta _{S_{k-1}}$ holds, we also have
1818: $\d(X_{S_{k+1}}, X_{S_k}) \geq \eps_3/8$ for the same range of $k$ (to
1819: see this, recall from Step 1(e) that $\eps_6 < \eps_3/8$). Hence
1820: $X_{S_{2(j+1)N-2}}$ must be outside $\ol D$, according to the
1821: definition of $\eps_4$ and $N$ in Step 1(d). Since $X$ always
1822: stays inside $\ol D$, at least one of the events $F_6(S_1) \circ
1823: \theta _{S_{k-1}}$ must fail for some $2jN \leq k \leq 2(j+1)N
1824: -1$. Hence, at least one event $F_1(S_1) \circ \theta _{S_{k-1}}$
1825: holds for some $2jN \leq k \leq 2(j+1)N -1$. Since $F_4(S_1) \circ
1826: \theta _{S_{k-1}}$ holds for every $k \leq 2NN_0$, there is a
1827: reduction of at least $\eps_7/2$ in the distance between $X$ and
1828: $Y$ on every interval $[S_{2jN}, S_{2(j+1)N}]$, that is,
1829: $$\d \left( X_{S_{2jN}}, \ Y_{S_{2jN}} \right)
1830: \leq \d \left( X_{S_{2(j-1)N}}, \ Y_{S_{2(j-1)N}} \right)
1831: - {\eps_7\over 2} \qquad \hbox{for every } j\in \{ 1, \cdots, N_0 \}.
1832: $$
1833: Summing over $j$ we obtain
1834: $$ \d \left( X_{S_{2N_0N}}, \ Y_{S_{2N_0N}} \right) \leq \d( X_0, \ Y_0 )
1835: - {N_0 \eps_7\over 2} \leq \d(X_0, Y_0)-2\rho <0.
1836: $$
1837: This contradiction proves our claim that
1838: $$
1839: \hbox{if } \ \bigcap_{k\leq 2NN_0} \{ S_k-S_{k-1} \leq t_1\} \
1840: \hbox{holds, then } \ \d(X_t, Y_t) \leq \eps_1 \ \hbox{ for some
1841: } \ t \leq 2t_1 NN_0 .
1842: $$
1843:
1844: We have shown that $\d(X_t,Y_t) \leq \eps_1$ for some $t\leq 2t_1
1845: NN_0$ with probability greater than $p_2 \df p_1^{2NN_0}>0$. By
1846: the Markov property, $\d(X_t,Y_t) \leq \eps_1$ for some $t\leq
1847: 2jt_1 NN_0$ with probability greater than $1- (1-p_2)^j$. To
1848: complete the proof, it suffices to let $j\to \infty$.
1849: \qed
1850:
1851:
1852:
1853: \bigskip
1854:
1855: Two of the assumptions on the boundary of $D$ made in Theorem 4.3,
1856: that it is convex with $C^1$-smooth boundary and it does not
1857: contain any line segments, are convenient from the technical point
1858: of view but most likely one can dispose of them with analysis more
1859: refined than that in our proof.
1860:
1861:
1862: \bigskip
1863: \noindent{\bf Example 4.4}. Suppose that $D$ is the annulus
1864: $\{x\in \R^2: 1<|x|<2\}$. The rotation of $D$ around $(0,0)$ with
1865: an angle in $(0, 2\pi)$ is an isometry with no fixed points.
1866: Hence, there exists a shy coupling of reflected Brownian motions
1867: in this annulus (see Example 3.3).
1868:
1869: \bigskip
1870:
1871: There are many open problems concerning existence of shy couplings
1872: but we find the following two questions especially intriguing.
1873: Recall that $\cB(x,r)$ denotes the open ball with center $x$ and
1874: radius $r$.
1875:
1876: \bigskip
1877: \noindent{\bf Open problems 4.5}. {\sl (i) Does there exist a shy
1878: coupling of reflected Brownian motions in $\cB((0,0),3) \setminus
1879: \cB((1,0),1)$?
1880:
1881: (ii) Does there exist a shy coupling of reflected Brownian motions
1882: in any simply connected planar domain?
1883:
1884: }
1885:
1886: \bigskip
1887: We end this paper with a vague remark concerning a potential
1888: relationship between shy couplings and an old and well known
1889: problem of ``fixed points.'' Suppose that $\cS$ is a topological
1890: space. If every continuous mapping $I: \cS \to \cS$ has a fixed
1891: point, i.e., a point $x\in\cS$ such that $I(x) = x$, then we say
1892: that $\cS$ has the fixed point property. One of the most famous
1893: fixed point theorems is that of Brouwer---it asserts that a closed
1894: ball in $\R^d$ has the fixed point property. Spheres obviously do
1895: not have the fixed point property. Some of our results may suggest
1896: that a shy coupling exists if and only if the state space does not
1897: have the fixed point property. Example 3.8 applied to the graph
1898: illustrated in Fig.~3.6 shows that this conjecture is false at
1899: this level generality. It is possible, though, that a weaker form
1900: of this assertion is true---we leave it as an open problem.
1901:
1902:
1903: \vskip1truein
1904:
1905:
1906: \centerline{REFERENCES}
1907: \bigskip
1908:
1909: \item{[BC1]} K.~Burdzy and Z.-Q.~Chen, Local time flow
1910: related to skew Brownian motion. {\it Ann. Probab. \bf 29} (2001),
1911: 1693-1715.
1912:
1913: \item{[BC2]} K.~Burdzy and Z.-Q.~Chen, Coalescence of
1914: synchronous couplings. {\it Probab. Theory Rel. Fields \bf 123}
1915: (2002), 553--578.
1916:
1917: \item{[BCJ]} K.~Burdzy, Z.-Q.~Chen and P.~Jones, Synchronous
1918: couplings of reflected Brownian motions in smooth domains.
1919: Preprint, 2005.
1920:
1921: \item{[BK]} K.~Burdzy and H.~Kaspi, Lenses in skew Brownian
1922: flow. {\it Ann. Probab. \bf 32} (2004), 3085--3115.
1923:
1924: \item{[FW]} M.~Freidlin and A.~Wentzell, Diffusion processes
1925: on graphs and the averaging principle. {\it Ann. Probab. \bf 21}
1926: (1993), 2215-2245.
1927:
1928: \item{[F]} A.~Friedman, {\it Differential Games}. Wiley, New
1929: York, 1971.
1930:
1931: \item{[HS]} J.M.~Harrison and L.A.~Shepp, On skew Brownian
1932: motion. {\it Ann. Probab}. {\bf 9} (1981), 309-313.
1933:
1934: \item{[KS]} I.~Karatzas and S. E.~Shreve, {\it Brownian Motion and
1935: Stochastic Calculus}, Second edition. Springer, New York, 1991.
1936:
1937: \item{[L]} T.~Lindvall, {\it Lectures on the Coupling
1938: Method}. Wiley, New York, 1992.
1939:
1940: \item{[LS]} P.~L. Lions and A.~S. Sznitman, Stochastic
1941: differential equations with reflecting boundary conditions. {\it
1942: Comm. Pure Appl. Math.} {\bf 37} (1984), 511-537.
1943:
1944: \item{[RY]} D.~Revuz and M.~Yor, {\it Continuous Martingales
1945: and Brownian Motion}. Springer, New York, 1991.
1946:
1947: \item{[Ta]} H. Tanaka, Stochastic differential equations with
1948: reflecting boundary condition in convex regions. {\it Hiroshima
1949: Math. J. \bf 9} (1979), 163-177.
1950:
1951:
1952: \vskip1truein
1953:
1954:
1955: \bigskip
1956: \noindent I.B.: Department of Mathematics, Weizmann Institute of
1957: Science, Rehovot 76100, Israel {\tt itai.benjamini@weizmann.ac.il}
1958:
1959:
1960: \bigskip
1961: \noindent K.B. and Z.C.: Department of Mathematics, Box 354350,
1962: University of Washington, Seattle, WA 98115-4350, USA \hfill\break
1963: {\tt burdzy@math.washington.edu, zchen@math.washington.edu}
1964:
1965:
1966:
1967: \bye
1968: