This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the duration of stays of Brownian motion in domains in Euclidean space

Dimitrios Betsakos D. Betsakos: Department of Mathematics, Aristotle University of Thessaloniki, GR-54124 Thessaloniki, Greece. betsakos@math.auth.gr Maher Boudabra M. Boudabra: Department of Mathematics, Monash University, Australia maher.boudabra@monash.edu  and  Greg Markowsky G. Markowsky: Department of Mathematics, Monash University, Australia greg.markowsky@monash.edu
(Date: May 2021)
Abstract.

Let TDT_{D} denote the first exit time of a Brownian motion from a domain DD in n{\mathbb{R}}^{n}. Given domains U,WnU,W\subseteq{\mathbb{R}}^{n} containing the origin, we investigate the cases in which we are more likely to have fast exits from UU than WW, meaning 𝐏(TU<t)>𝐏(TW<t){\bf P}(T_{U}<t)>{\bf P}(T_{W}<t) for tt small. We show that the primary factor in the probability of fast exits from domains is the proximity of the closest regular part of the boundary to the origin. We also prove a result on the complementary question of longs stays, meaning 𝐏(TU>t)>𝐏(TW>t){\bf P}(T_{U}>t)>{\bf P}(T_{W}>t) for tt large. This result, which applies only in two dimensions, shows that the unit disk has the lowest probability of long stays amongst all Schlicht domains.

Key words and phrases:
Brownian motion, exit time distribution, capacity
1991 Mathematics Subject Classification:
60J65, 60J45

1. Introduction

The distribution of the exit time of a Brownian motion from a domain in n{\mathbb{R}}^{n} gives a measure of the size and information about the shape of the domain. Naturally, a small domain will have an exit time which is smaller, in some sense, than that of a large domain. However, making this statement precise is a bit tricky, and this question has been of interest to a number of researchers, especially in two dimensions. For examples, the reader is referred to the recent papers [2, 3, 4, 15, 18], as well as to the older works [9, 11]. This paper is concerned with this question, and particularly with the relationship between the probability that the exit time is small and the proximity of the boundary of the domain to the starting point of the Brownian motion.

Before stating our results, let us fix notation. Let 𝐁t{\bf B}_{t}, t0t\geq 0 denote standard Brownian motion moving in n{\mathbb{R}}^{n}, n2n\geq 2. We denote by 𝐏{\bf P} and 𝐄{\bf E} the corresponding probability measure and expectation, respectively, and will use superscripts in 𝐏x{\bf P}^{x} and 𝐄x{\bf E}^{x} to signify that we are using the probability measure associated with Brownian motion starting from xx (see e.g. [21]). If ever the starting point is not mentioned, then we assume it is the origin. For a domain DD containing the origin in n{\mathbb{R}}^{n}, we denote by TDT_{D} the first exit time of 𝐁t{\bf B}_{t} from DD; that is

TD=inf{t0:𝐁tD}.T_{D}=\inf\{t\geq 0:{\bf B}_{t}\notin D\}.

We also let d(A){\rm d}(A) denote the distance from the origin to the set AnA\subset{\mathbb{R}}^{n}; that is,

d(A)=inf{x:xA}.{\rm d}(A)=\inf\{\|x\|:x\in A\}.

For our first result, we consider the quantity 𝐏(TD<t){\bf P}(T_{D}<t) for tt small (this is what we mean by “fast exits”) and its relation to d(D){\rm d}(\partial D). The authors of this paper have already proved some results in this direction, in [7]. There, we considered this problem only in two dimensions, and assumed also that the domains in question were simply connected. Our main result in that paper was the following.

Theorem A. [7] Suppose that U,WU,W are two simply connected domains in 2{\mathbb{R}}^{2} both containing the origin, and that d(U)<d(W){\rm d}(\partial U)<{\rm d}(\partial W). Then, for all sufficiently small t>0t>0,

(1.1) 𝐏0(TU<t)>𝐏0(TW<t).{\bf P}^{0}(T_{U}<t)>{\bf P}^{0}(T_{W}<t).

In fact,

limt0+𝐏(TU<t)𝐏(TW<t)=.\lim_{t\to 0^{+}}\frac{{\bf P}(T_{U}<t)}{{\bf P}(T_{W}<t)}=\infty.

The proof there depended heavily upon simple connectivity and the topology of the plane, and did not seem at the time to generalize to higher dimensions. However, subsequent study has shown that the result does admit a considerable generalization to n{\mathbb{R}}^{n}, and furthermore that the requirement of simple connectivity was unnecessary. Dropping the requirement of simple connectivity, however, does significantly complicate the proof, and compels us to introduce the concept of regular boundary points, which we now discuss.

For a closed set KnK\subset{\mathbb{R}}^{n}, let τK=inf{t>0:BtK}\tau_{K}=\inf\{t>0:B_{t}\in K\}. We will call this the hitting time of the set KK. A point xx is called regular for KK if 𝐏x(τK=0)=1{\bf{P}}^{x}(\tau_{K}=0)=1 and irregular otherwise (see [21] or [24]). Intuitively speaking, a regular point xx for KK is a point from which the Brownian motion hits K\{x}K\backslash\{x\} immediately upon starting at xx. Note that 𝐏x(τK=0){0,1}{\bf{P}}^{x}(\tau_{K}=0)\in\{0,1\} by Blumenthal’s zero-one law; that is, if xx is irregular then 𝐏x(τK=0)=0{\bf{P}}^{x}(\tau_{K}=0)=0. Let KrK^{\rm r} denote the set of regular points of KK. This set contains the interior of KK . This leaves only points in K\partial K as unknowns, and these may or may not be regular. A simple example of a planar domain with an irregular boundary point is 𝔻\{0}{\mathbb{D}}\backslash\{0\}, and the boundary point 0 is irregular. On the other hand, it is well known that all boundary points in any simply connected domain in the plane (strictly smaller than {\mathbb{C}} itself) are regular (see [22], for instance).

The main result of this paper can now be stated, and is as follows.

Theorem 1.

Let U,WU,W be domains in n{\mathbb{R}}^{n}, both containing the origin. Suppose that d((U)r)<d((W)r){\rm d}((\partial U)^{r})<{\rm d}((\partial W)^{r}). Then

(1.2) limt0+𝐏0(TU<t)𝐏0(TW<t)=.\lim_{t\to 0+}\frac{{\bf P}^{0}(T_{U}<t)}{{\bf P}^{0}(T_{W}<t)}=\infty.

We also consider the complementary problem of “long stays”; that is, the behavior of the quantity 𝐏(TD<t){\bf P}(T_{D}<t) for tt large. For this problem, we will focus only on the two-dimensional case. Again, we have a previous result on this topic in our previous paper [7]. To describe this result, we first need a definition. For a domain D2D\subset{\mathbb{R}}^{2}, let

H(D)=sup{p>0:𝐄[(TD)p]<};{\rm H}(D)=\sup\{p>0:{\bf E}[(T_{D})^{p}]<\infty\};

note that H(D){\rm H}(D) is proved in [9] to be exactly equal to half of the Hardy number of DD, a purely analytic quantity, as defined in [13], and is therefore calculable for a number of common domains. The domains we consider satisfy a normalization condition: they are Schlicht. A planar simply connected doamin DD\subsetneqq{\mathbb{C}} is Schlicht if it contains the origin and D=f(𝔻)D=f({\mathbb{D}}), where ff is the Riemann map with f(0)=0f(0)=0 and f(0)=1f^{\prime}(0)=1. It is known that H(D)14{\rm H}(D)\geq\frac{1}{4} as long as DD\neq{\mathbb{C}} is Schlicht ([9]). The following is our prior result.

Theorem B. [7] Suppose that U,WU,W are Schlicht domains and H(U)>H(W){\rm H}(U)>{\rm H}(W). Then

(1.3) lim supt𝐏(TW>t)𝐏(TU>t)=.\limsup_{t\to\infty}\frac{{\bf P}(T_{W}>t)}{{\bf P}(T_{U}>t)}=\infty.

We conjectured there that this proposition remained true with the lim sup\limsup replaced by lim\lim, but we were not able to prove it (except when WW is a wedge). Subsequent study has revealed another result in this direction. Before stating the result, we provide a bit of motivation.

In [11], Davis explored the relation of planar Brownian motion to classical complex analysis (for more on this topic, see [16] and the references therein). Among a number of important ideas was Davis’ statement that “the distribution of TDT_{D} is an intuitively appealing measure of the size of DD” (notation changed to match that in this paper). Davis then suggested applying this idea to the set of Schlicht domains. He conjectured that if DD is a Schlicht domain, then

𝐏(TD<t)𝐏(T𝔻<t),{\bf P}(T_{D}<t)\leq{\bf P}(T_{\mathbb{D}}<t),

for all t>0t>0. However, McConnell disproved this for sufficiently small tt and DD an infinite strip, in [17]; note that this follows also from our results on fast exits (Theorem A), since d(D)<1{\rm d}(\partial D)<1 for any Schlicht domain DD other than 𝔻{\mathbb{D}}, as a consequence of Schwarz’s lemma. We will prove that Davis’ conjecture is correct for large tt, i.e. for long stays. Our result is as follows.

Theorem 2.

Let DD be a Schlicht domain other than 𝔻{\mathbb{D}}. Then there exists to>0t_{o}>0 such that for every t>tot>t_{o},

(1.4) 𝐏0(TD<t)<𝐏0(T𝔻<t).{\bf P}^{0}(T_{D}<t)<{\bf P}^{0}(T_{\mathbb{D}}<t).

To prove Theorems 1 and 2, we will use a number of known deep results. The next section contains the necessary preliminaries, and the subsequent section contains the proofs. A short final section contains some concluding remarks.

2. Preliminaries

In this section, we collect the topics and results we will need for the proofs of our theorems.

2.1. Killed Brownian motion

A Brownian motion running in n{\mathbb{R}}^{n} admits the transition density

p(t,x,y)=1(2πt)n/2e|xy|22t.p(t,x,y)=\frac{1}{(2\pi t)^{n/2}}e^{\frac{-|x-y|^{2}}{2t}}.

This means that the probability that Brownian motion starting at xx is in a Borel set AA at time tt is equal to Ap(t,x,y)𝑑y\int_{A}p(t,x,y)dy. Note that pp satisfies the heat equation: tp=12Δp\partial_{t}p=\frac{1}{2}\Delta p. Our primary interest will be in the “stopped transition density” pD(t,x,y)p_{D}(t,x,y) (see e.g. [21]), which is taken in relation to a domain DD in n{\mathbb{R}}^{n} and applies to Browian motion killed upon leaving DD. In particular, if KK is outside DD, then KpD(t,x,y)𝑑y=0\int_{K}p_{D}(t,x,y)dy=0. By the strong Markov property, the formula of pD(t,x,y)p_{D}(t,x,y) is given by

pD(t,x,y)=p(t,x,y)𝐄x(p(tTD,BTD,y)1TD<t)p_{D}(t,x,y)=p(t,x,y)-{\bf E}^{x}(p(t-T_{D},B_{T_{D}},y)1_{T_{D}<t})

and one can see that pD(t,x,y)p(t,x,y)p_{D}(t,x,y)\leq p(t,x,y). Note that for every Borel set KnK\subset{\mathbb{R}}^{n}, the identity

𝐏x(BtK,t<TD)=KpD(t,x,y)mn(dy){\bf P}^{x}(B_{t}\in K,t<T_{D})=\int_{K}p_{D}(t,x,y)\,m_{n}(dy)

persists; here and below mnm_{n} is the nn-dimensional Lebesgue measure. Furthermore, pDp_{D} also satisfies the heat equation in DD with zero boundary values.

Two basic properties of the transition density pDp_{D} are the domain monotonicity: if DΩD\subset\Omega then pD(t,x,y)pΩ(t,x,y)p_{D}(t,x,y)\leq p_{\Omega}(t,x,y), and the semigroup property:

pD(t+s,x,y)=DpD(t,x,z)pD(s,z,y)mn(dz),t,s>0,x,yD.p_{D}(t+s,x,y)=\int_{D}p_{D}(t,x,z)p_{D}(s,z,y)\;m_{n}(dz),\;\;\;t,s>0,\;x,y\in D.

An additional property of pDp_{D} is the principle of “not feeling the boundary”. We will need this principle in its simplest form (see [10]): If DD is a convex domain in n{\mathbb{R}}^{n}, then for every x,yDx,y\in D,

limt0+pD(t,x,y)p(t,x,y)=1.\lim_{t\to 0+}\frac{p_{D}(t,x,y)}{p(t,x,y)}=1.

When DD is a ball, the stopped transition density has a radial monotonicity property. Probably, this property is known to experts, but we haven’t found it in the literature, and thus we provide a proof.

Lemma 1.

Let DD be the ball in n{\mathbb{R}}^{n} centered at the origin with radius RR. Then for every t>0t>0, the function pD(t,0,x)p_{D}(t,0,x) is radially strictly decreasing.

Proof.
Refer to caption
Figure 1. The ball, D+D_{+}, and DD_{-}.

By symmetry, pD(t,0,x)p_{D}(t,0,x) is a radial function. Let e1=(1,0,,0)e_{1}=(1,0,\dots,0) be the first coordinate vector and let 0<r<s<R0<r<s<R. Consider the (n1)(n-1)-dimensional plane PP (a line when n=2n=2), perpendicular to e1e_{1} and passing from the point r+s2e1\frac{r+s}{2}\,e_{1}. Let

D+:={x=(x1,x2,,xn)D:x1>r+s2},D_{+}:=\{x=(x_{1},x_{2},\dots,x_{n})\in D:x_{1}>\frac{r+s}{2}\},
D:={x=(x1,x2,,xn)D:x1<r+s2}.D_{-}:=\{x=(x_{1},x_{2},\dots,x_{n})\in D:x_{1}<\frac{r+s}{2}\}.

Note that the reflection of D+D_{+} in PP is contained in DD_{-}, and that the reflection of the point se1se_{1} in PP is the point re1re_{1}.

Now, since pDp_{D} satisfies the heat equation, we can use a special case of a polarization result [8, Theorem 9.4], [5, Theorem 4], and conclude that

(2.1) pD(t,0,se1)<pD(t,0,re1).p_{D}(t,0,se_{1})<p_{D}(t,0,re_{1}).

2.2. Green’s function

An important related object is Green’s function. We give its probabilistic definition; see e.g. [21], [24]. With pDp_{D} as above, we define

GD(x,y)=0pD(s,x,y)𝑑s.G_{D}(x,y)=\int_{0}^{\infty}p_{D}(s,x,y)ds.

2.3. Logarithmic and Newtonian capacity

We now discuss the capacity of a set in n{\mathbb{R}}^{n}. The definitions of capacity of sets in n{\mathbb{R}}^{n} are not entirely standardized in the literature; we will follow the definitions used in [21], [22]. For a compact set KK in n{\mathbb{R}}^{n} with n2n\geq 2, define

Rn(K):=infK×Kf(xy)μ(dx)μ(dy),R_{n}(K):=\inf\int_{K\times K}f(x-y)\mu(dx)\mu(dy),

where

f(x)={ln|x|,if n=2|x|2n,if n3,f(x)=\begin{cases}-\ln|x|,&\text{if }n=2\\ |x|^{2-n},&\text{if $n\geq 3$},\end{cases}

and where the infinimum is taken over all probability measures μ\mu having their support on KK. If K2K\subset{\mathbb{R}}^{2}, its logarithmic capacity is defined by c2(K)=eR2(K)c_{2}(K)=e^{-R_{2}(K)}. If KnK\subset{\mathbb{R}}^{n}, n3n\geq 3, its Newtonian capacity is cn(K)=Rn(K)1c_{n}(K)=R_{n}(K)^{-1}. These capacities have an important connection with Brownian motion in nn dimensions, as we now describe.

A compact set KK is referred to as polar when 𝐏(τK<)=0{\bf P}(\tau_{K}<\infty)=0, i.e when it is not hit by Brownian motion with probability one. Otherwise, it is called nonpolar. By a theorem of Kakutani (see e.g. [19, Prop. 6.1]), a compact set KnK\subset{\mathbb{R}}^{n} is nonpolar if and only if c2(K)>0c_{2}(K)>0.

It is worth mentioning that the 𝐏(τK<){\bf P}(\tau_{K}<\infty) can take values other than zero or one if the dimension nn is at least 3. To see this, consider the ball {x<r}\{\mid x\mid\ <r\}. Then the probability to hit that ball starting from x{x<r}x\not\in\{\mid x\mid\ <r\} is given by (rx)n2(\frac{r}{\mid x\mid})^{n-2} (see e.g. [21, p.56]), where n3n\geq 3 is the dimension of the space.

2.4. Condenser capacity

We now discuss a related topic, the capacity of a condenser. A condenser is a pair (D,K)(D,K), where DD is a region and KDK\subset D is a compact set, both in n{\mathbb{R}}^{n}. The capacity of (D,K)(D,K) is defined to be the infimum of the Dirichlet integral

cap(D,K)=G\D|u|2mn(dx),{\rm cap}(D,K)=\int_{G\backslash D}|\nabla u|^{2}m_{n}(dx),

where the infimum is taken over all smooth functions uu with u=1u=1 on KK and u=0u=0 on D\partial D. This quantity has a number of nice properties that will be relevant to us. For instance, Dirichlet’s principle implies that if a minimizer exists then it must be harmonic. When the minimizer does exist it can be interpreted in terms of Brownian motion by u(z)=𝐏z(BTD\KK)u(z)={\bf P}^{z}(B_{T_{D\backslash K}}\in K). Furthermore, this type of capacity in two dimensions can be shown to be conformally invariant, in the sense that if ff is a conformal map from DD onto another domain DD^{\prime}, then cap(D,K)=cap(D,f(K)){\rm cap}(D,K)={\rm cap}(D^{\prime},f(K)). The condenser capacity cap(D,K){\rm cap}(D,K) is also known as the Green capacity of KK with respect to DD. See [1], [25], [21] for more on this topic.

2.5. Equilibrium measure

Let KK be a compact subset of a domain DD in n{\mathbb{R}}^{n}. We assume that DD possesses a finite Green’s function GDG_{D}. Then there exists a unique Borel measure μK,D\mu_{K,D} on KK such that

(2.2) 𝐏x(τK<TD)=KGD(x,y)μK,D(dy),xD.{\bf P}^{x}(\tau_{K}<T_{D})=\int_{K}G_{D}(x,y)\,\mu_{K,D}(dy),\;\;\;x\in D.

This is the equilibrium measure of KK with respect to DD. Its total measure is equal to the capacity of the condenser (D,K)(D,K): μK,D(K)=cap(D,K)\mu_{K,D}(K)={\rm cap}(D,K); see [21, Chapter 6].

2.6. Fundamental frequency

If DD is a planar domain, its fundamental frequency is given by

(2.3) λ(D)=infϕD|ϕ|2𝑑mnDϕ2𝑑mn,\lambda(D)=\inf_{\phi}\frac{\int_{D}|\nabla\phi|^{2}\,dm_{n}}{\int_{D}\phi^{2}\,dm_{n}},

where the infimum is taken over all smooth functions ϕ\phi with compact support in DD. If the Laplacian has a sequence of Dirichlet eigenvalues on DD (e.g. when mn(D)<m_{n}(D)<\infty), then λ(D)\lambda(D) represents the first eigenvalue. We note, however, that λ(D)\lambda(D) is defined by (2.3) even when there are no eigenvalues, and that λ(D)\lambda(D) may be equal to zero.

The next theorem (see e.g. [24, §3.1]) gives the connection between the principal Dirichlet eigenvalue and Brownian motion which we will exploit.

Theorem C. If DD is a domain in \mathbb{C}, then for every zDz\in D,

(2.4) limt+2tlog1𝐏z(TD>t)=λ(D).\lim_{t\to+\infty}\frac{2}{t}\log\frac{1}{{\bf P}^{z}(T_{D}>t)}=\lambda(D).

The following theorem provides a characterization of the disk as an extremal representative of the class of Schlicht domains; recall that these are images of the unit disk under conformal maps ff with f(0)=0f(0)=0 and f(0)=1f^{\prime}(0)=1. For a proof, see [20, §5.8] and [14].

Theorem D. If DD is a Schlicht domain then λ(D)λ(𝔻)\lambda(D)\leq\lambda(\mathbb{D}), with equality if and only if D=𝔻D=\mathbb{D}.

Armed with this large assortment of tools, we can now tackle the proofs of our theorems.

3. Proofs of Theorems 1 and 2

The proof of the theorem differs in details depending on whether the dimension nn satisfies n=2n=2 or n3n\geq 3. We will give a complete proof in the case n=2n=2, and then indicate how the proof must be adjusted when n3n\geq 3. The heart of the proof is contained in the following lemma.

Lemma 2.

Let KK be a compact set in the plane with 0K0\notin K and c2(K)>0{\rm c}_{2}(K)>0. Let aa be a regular point of KK. Let δ>0\delta>0. There exist positive constants C=C(K,δ)C=C(K,\delta) and T=T(K,δ)T=T(K,\delta) such that for every t(0,T)t\in(0,T),

(3.1) 𝐏0(τK<t)Ce(|a|+δ)2/(2t).{\bf P}^{0}(\tau_{K}<t)\geq C\,e^{-(|a|+\delta)^{2}/(2t)}.
Proof.

We will use a modification of trick taken from [12, Proof of Lemma 3.6].

Set L:=KD(a,δ/5)¯L:=K\cap\overline{D(a,\delta/5)} and Ω:=D(0,3|a|+δ)\Omega:=D(0,3|a|+\delta). We will eventually stop the Brownian motion upon exiting Ω\Omega, which allows to use the equilibrium measure. Figure 2 may help the reader understand this setup.

Refer to caption
Figure 2. K,L,K,L, and Ω\Omega.

Note that aa is a regular point of LL and c2(L)>0c_{2}(L)>0. Since LKL\subset K, for every t>0t>0 we have

(3.2) 𝐏0(τL<t)𝐏0(τK<t)\displaystyle{\bf P}^{0}(\tau_{L}<t)\leq{\bf P}^{0}(\tau_{K}<t)

Observe also that for t>0t>0,

𝐏0(τL<t)\displaystyle{\bf P}^{0}(\tau_{L}<t) \displaystyle\geq 𝐏0(τL<tandτL<TΩ)\displaystyle{\bf P}^{0}(\tau_{L}<t\;\;\;\hbox{and}\;\;\;\tau_{L}<T_{\Omega})
=\displaystyle= 𝐏0(τL<TΩ)𝐏0(τLtandτL<TΩ).\displaystyle{\bf P}^{0}(\tau_{L}<T_{\Omega})-{\bf P}^{0}(\tau_{L}\geq t\;\;\;\hbox{and}\;\;\;\tau_{L}<T_{\Omega}).

By [21, Theorem 5.1, p. 190], for every zz\in{\mathbb{C}},

𝐏z(τL<TΩ)\displaystyle{\bf P}^{z}(\tau_{L}<T_{\Omega}) =\displaystyle= LGΩ(z,x)μL,Ω(dx)\displaystyle\int_{L}G_{\Omega}(z,x)\mu_{L,\Omega}(dx)
=\displaystyle= L0pΩ(s,z,x)𝑑sμL,Ω(dx),\displaystyle\int_{L}\int_{0}^{\infty}p_{\Omega}(s,z,x)\,ds\,\mu_{L,\Omega}(dx),

where GΩG_{\Omega} is the Green function for Ω\Omega, pΩp_{\Omega} is the transition density for Brownian motion killed upon exiting Ω\Omega and μL,Ω\mu_{L,\Omega} is the equilibrium measure of LL with respect to Ω\Omega.

By the (simple) Markov property, the equation (3), Fubini’s theorem, the domain monotonicity, the semigroup property of the transition density, and a change of variables,

𝐏0(τLtandτL<TΩ)\displaystyle{\bf P}^{0}(\tau_{L}\geq t\;\;\;\hbox{and}\;\;\;\tau_{L}<T_{\Omega})
=\displaystyle= ΩLpΩL(t,0,z)𝐏z(τL<TΩ)mn(dz)\displaystyle\int_{\Omega\setminus L}p_{\Omega\setminus L}(t,0,z)\;{\bf P}^{z}(\tau_{L}<T_{\Omega})\;m_{n}(dz)
=\displaystyle= ΩLpΩL(t,0,z)L0pΩ(s,z,x)𝑑sμL,Ω(dx)mn(dz)\displaystyle\int_{\Omega\setminus L}p_{\Omega\setminus L}(t,0,z)\int_{L}\int_{0}^{\infty}p_{\Omega}(s,z,x)ds\;\mu_{L,\Omega}(dx)\;m_{n}(dz)
=\displaystyle= L0ΩLpΩL(t,0,z)pΩ(s,z,x)mn(dz)𝑑sμL,Ω(dx)\displaystyle\int_{L}\int_{0}^{\infty}\int_{\Omega\setminus L}p_{\Omega\setminus L}(t,0,z)p_{\Omega}(s,z,x)m_{n}(dz)\;ds\;\mu_{L,\Omega}(dx)
\displaystyle\leq L0ΩpΩ(t,0,z)pΩ(s,z,x)mn(dz)𝑑sμL,Ω(dx)\displaystyle\int_{L}\int_{0}^{\infty}\int_{\Omega}p_{\Omega}(t,0,z)p_{\Omega}(s,z,x)m_{n}(dz)\;ds\;\mu_{L,\Omega}(dx)
=\displaystyle= L0pΩ(t+s,0,x)𝑑sμL,Ω(dx)\displaystyle\int_{L}\int_{0}^{\infty}p_{\Omega}(t+s,0,x)\,ds\;\mu_{L,\Omega}(dx)
=\displaystyle= LtpΩ(s,0,x)𝑑sμL,Ω(dx).\displaystyle\int_{L}\int_{t}^{\infty}p_{\Omega}(s,0,x)ds\;\mu_{L,\Omega}(dx).

Combining (3.2), (3), (3), (3), we obtain

𝐏0(τK<t)\displaystyle{\bf P}^{0}(\tau_{K}<t)
\displaystyle\geq L0pΩ(s,0,x)𝑑sμL,Ω(dx)LtpΩ(s,0,x)𝑑sμL,Ω(dx)\displaystyle\int_{L}\int_{0}^{\infty}p_{\Omega}(s,0,x)ds\;\mu_{L,\Omega}(dx)-\int_{L}\int_{t}^{\infty}p_{\Omega}(s,0,x)ds\;\mu_{L,\Omega}(dx)
=\displaystyle= L0tpΩ(s,0,x)𝑑sμL,Ω(dx).\displaystyle\int_{L}\int_{0}^{t}p_{\Omega}(s,0,x)ds\;\mu_{L,\Omega}(dx).

We will use the fact that for the disk Ω\Omega, the transition density pΩ(s,0,x)p_{\Omega}(s,0,x) is a decreasing function of |x||x| (Lemma 1). We will also use the fact that μL,Ω(L)\mu_{L,\Omega}(L) is equal to the Green capacity (or condenser capacity) cap(Ω,L){\rm cap}(\Omega,L), and obtain:

(3.7) L0tpΩ(s,0,x)𝑑sμL,Ω(dx)\displaystyle\int_{L}\int_{0}^{t}p_{\Omega}(s,0,x)ds\;\mu_{L,\Omega}(dx) \displaystyle\geq L0tpΩ(s,0,|a|+δ/5)𝑑sμL,Ω(dx)\displaystyle\int_{L}\int_{0}^{t}p_{\Omega}(s,0,|a|+\delta/5)ds\;\mu_{L,\Omega}(dx)
=\displaystyle= cap(Ω,K)0tpΩ(s,0,|a|+δ/5)𝑑s.\displaystyle{\rm cap}(\Omega,K)\int_{0}^{t}p_{\Omega}(s,0,|a|+\delta/5)\,ds.

Now we use “the principle of not feeling the boundary” (see Subsection 2.1) and find a positive T1=T1(K,δ)T_{1}=T_{1}(K,\delta) such that

(3.8) pΩ(s,0,|a|+δ/5)12p(s,0,|a|+δ/5),s(0,T1).p_{\Omega}(s,0,|a|+\delta/5)\geq\frac{1}{2}p(s,0,|a|+\delta/5),\;\;\;\;\;s\in(0,T_{1}).

By (3), (3.7), and (3.8), for t(0,T1)t\in(0,T_{1}),

𝐏0(τK<t)\displaystyle{\bf P}^{0}(\tau_{K}<t) \displaystyle\geq cap(Ω,L)20tp(s,0,|a|+δ/5)𝑑s\displaystyle\frac{{\rm cap}(\Omega,L)}{2}\;\int_{0}^{t}p(s,0,|a|+\delta/5)ds
=\displaystyle= cap(Ω,L)20t12πse(|a|+δ/5)2/(2s)𝑑s.\displaystyle\frac{{\rm cap}(\Omega,L)}{2}\;\int_{0}^{t}\frac{1}{2\pi s}e^{-(|a|+\delta/5)^{2}/(2s)}\;ds.

By elementary calculus, there exists a positive number T<T1T<T_{1} such that for every t(0,T)t\in(0,T),

(3.10) 0t12πse(|a|+δ/5)2/(2s)𝑑s12e(|a|+δ)2/(2t).\int_{0}^{t}\frac{1}{2\pi s}e^{-(|a|+\delta/5)^{2}/(2s)}\;ds\geq\frac{1}{2}e^{-(|a|+\delta)^{2}/(2t)}.

Also, by [6, Lemma 1], and denoting by LL^{*} the disk D(0,c2(L))D(0,c_{2}(L)), we get

(3.11) cap(Ω,L)cap(Ω,L¯)=(log3|a|+δc2(L))1,{\rm cap}(\Omega,L)\geq{\rm cap}(\Omega,\overline{L^{*}})=\left(\log\frac{3|a|+\delta}{{\rm c}_{2}(L)}\right)^{-1},

which is a quantity that depends only on KK and δ\delta.

We combine (3), (3.10), (3.11) to conclude that for t(0,T)t\in(0,T),

(3.12) 𝐏0(τK<t)Ce(|a|+δ)2/(2t).\displaystyle{\bf P}^{0}(\tau_{K}<t)\geq C\;e^{-(|a|+\delta)^{2}/(2t)}.


Proof of Theorem 1

We start with some reductions. First, we assume without loss of generality that d((W)r)=1{\rm d}((\partial W)^{r})=1. Then for every t>0t>0,

(3.13) 𝐏0(TW<t)𝐏0(T𝔻<t),{\bf P}^{0}(T_{W}<t)\leq{\bf P}^{0}(T_{\mathbb{D}}<t),

and so we may assume that W=𝔻W={\mathbb{D}}.

Let a(U)ra\in(\partial U)^{r} be a point with |a|<1|a|<1. Choose a disk D(a,r)D(a,r) with 0<r<1|a|0<r<1-|a|, and set K:=(U)D(a,r)¯K:=({\mathbb{C}}\setminus U)\cap\overline{D(a,r)}. Note that KK has positive logarithmic capacity and that UKU\subset{\mathbb{C}}\setminus K. Therefore,

(3.14) 𝐏0(TU<t)𝐏0(TK<t)=𝐏0(τK<t).{\bf P}^{0}(T_{U}<t)\geq{\bf P}^{0}(T_{{\mathbb{C}}\setminus K}<t)={\bf P}^{0}(\tau_{K}<t).

So we may assume that U=KU={\mathbb{C}}\setminus K, where KK is a compact subset of 𝔻{\mathbb{D}} with c2(K)>0{\rm c}_{2}(K)>0.

By Lemma 2, for every δ>0\delta>0, there exist positive constants C,TC,T depending on KK and δ\delta such that for every t(0,T)t\in(0,T),

(3.15) 𝐏0(τK<t)Ce(|a|+δ)2/(2t),{\bf P}^{0}(\tau_{K}<t)\geq C\;e^{-(|a|+\delta)^{2}/(2t)},

where aa is a regular point of KK. By taking δ\delta small enough, we may assume that |a|+δ<1|a|+\delta<1. On the other hand, an ingenious argument due to McConnell [17] shows that, for all t>0t>0 and all positive integers m3m\geq 3,

(3.16) 𝐏(T𝔻<t)c(m)ecos2(π/m)2t.{\bf P}(T_{\mathbb{D}}<t)\leq c(m)\;e^{-\frac{\cos^{2}(\pi/m)}{2t}}.

By fixing mm large enough, we see that for any ϵ>0\epsilon>0 we can find a constant Cϵ>0C_{\epsilon}>0 such that for all t>0t>0 we have

𝐏(T𝔻t)<Cϵe(1ϵ)2t.{\bf P}(T_{\mathbb{D}}\leq t)<C_{\epsilon}\;e^{-\frac{(1-\epsilon)}{2t}}.

Thus Lemma 2 gives

limt0+𝐏(TU<t)𝐏(TW<t)\displaystyle\lim_{t\to 0+}\frac{{\bf P}(T_{U}<t)}{{\bf P}(T_{W}<t)} \displaystyle\geq limt0+𝐏(τK<t)𝐏(T𝔻<t)\displaystyle\lim_{t\to 0+}\frac{{\bf P}(\tau_{K}<t)}{{\bf P}(T_{\mathbb{D}}<t)}
\displaystyle\geq limt0+C(K,δ)exp[(|a|+δ)22t]Cϵexp[1ϵ2t].\displaystyle\lim_{t\to 0+}\frac{C(K,\delta)\;\exp\left[-\frac{(|a|+\delta)^{2}}{2t}\right]}{C_{\epsilon}\;\exp\left[-\frac{1-\epsilon}{2t}\right]}.

Choose ϵ\epsilon and δ\delta small enough so that (|a|+δ)2<1ϵ(|a|+\delta)^{2}<1-\epsilon. Then the limit above is equal to \infty and this completes the proof of the result in two dimensions.

The proof in three or higher dimensions follows along exactly the same lines, with several necessary modifications, which we now indicate. In n3n\geq 3 dimensions, we will use the Newtonian capacity rather than the logarithmic. All other concepts used carry over directly to the higher dimensions with no real change, except that some of the estimates have to be changed. In particular, inequality (3.11) is specific to two dimensions; it is sufficient to replace it with [25, Cor. 1]. Furthermore, McConnell’s estimate (3.16) is also specific to two dimensions, but it can be replaced by [23, Cor. 3.4]. The result follows then as before. ∎


Proof of Theorem 2

Suppose that DD is a Schlicht domain other than 𝔻\mathbb{D}. By Theorems C and D (in Subsection 2.6),

(3.18) limt+(𝐏0(TD>t)𝐏0(T𝔻>t))2/t\displaystyle\lim_{t\to+\infty}\left(\frac{{\bf P}^{0}(T_{D}>t)}{{\bf P}^{0}(T_{\mathbb{D}}>t)}\right)^{2/t} =\displaystyle= limt+exp[2tlog𝐏0(TD>t)𝐏0(T𝔻>t)]\displaystyle\lim_{t\to+\infty}\exp\left[\frac{2}{t}\log\frac{{\bf P}^{0}(T_{D}>t)}{{\bf P}^{0}(T_{\mathbb{D}}>t)}\right]
=\displaystyle= exp(λ(𝔻)λ(D))>1.\displaystyle\exp(\lambda(\mathbb{D})-\lambda(D))>1.

It follows that there exists to>0t_{o}>0 such that for every t>tot>t_{o},

𝐏0(TD>t)>𝐏0(T𝔻>t),{\bf P}^{0}(T_{D}>t)>{\bf P}^{0}(T_{\mathbb{D}}>t),

which is equivalent to (1.4). ∎

References

  • [1] G.D. Anderson, M.K. Vamanamurthy,  The Newtonian capacity of a space condenser. Indiana Univ. Math. J. 34 (1985), no. 4, 753-776.
  • [2] R. Bañuelos, T. Carroll,   Brownian motion and the fundamental frequency of a drum, Duke Mathematical Journal, 75(1994), 575–602.
  • [3] R. Bañuelos, T. Carroll,   The maximal expected lifetime of Brownian motion, Mathematical Proceedings of the Royal Irish Academy, 2011, 1-11.
  • [4] R. Bañuelos, P. Mariano, J. Wang,   Bounds for exit times of Brownian motion and the first Dirichlet eigenvalue for the Laplacian, arxiv:2003.06867, 2020.
  • [5] D. Betsakos,  Equality cases in the symmetrization inequalities for Brownian transition functions and Dirichlet heat kernels. Ann. Acad. Sci. Fenn. Math. 33 (2008), no. 2, 413-427.
  • [6] D. Betsakos,  Geometric versions of Schwarz’s lemma for quasiregular mappings. Proc. Amer. Math. Soc. 139 (2011), no. 4, 1397-1407.
  • [7] D. Betsakos, M. Boudabra, G. Markowsky,  On the probability of fast exits and long stays of a planar Brownian motion in simply connected domains. J. Math. Anal. Appl. 493 (2021).
  • [8] F. Brock, A.Yu. Solynin,  An approach to symmetrization via polarization. Trans. Amer. Math. Soc. 352 (2000), 1759-1796.
  • [9] D. Burkholder,  Exit times of Brownian motion, harmonic majorization, and Hardy spaces. Advances in Math. 26 (1977), no. 2, 182-205.
  • [10] Z. Ciesielski,   Heat conduction and the principle of not feeling the boundary. Bull. Acad. Polon. Sci. Sér. Sci. Math. Astronom. Phys. 14 (1966), 435-440.
  • [11] B. Davis,  Brownian motion and analytic functions. Ann. Probab. 7 (1979), 913-932.
  • [12] A. Grigor’yan, L. Saloff-Coste,  Hitting probabilities for Brownian motion on Riemannian manifolds. J. Math. Pures Appl. (9) 81 (2002), no. 2, 115-142.
  • [13] L. Hansen,  Hardy classes and ranges of functions. Michigan Math. J. 17 (1970), 235-248.
  • [14] J. Hersch, On symmetric membranes and conformal radius: Some complements to Pólya’s and Szegö’s inequalitites. Arch. Rational Mech. Anal. 20 (1965), 378-390.
  • [15] D. Kim,   Quantitative inequalities for the expected lifetime of Brownian motion, Michigan Math. J., (to appear).
  • [16] G. Markowsky,  Planar Brownian Motion and Complex Analysis, arXiv:2012.08574, 2020.
  • [17] T.R. McConnell,  The size of an analytic function as measured by Lévy’s time change. Ann. Probab. 13 (1985), 1003-1005.
  • [18] J. P. Mendez-Hernández,   Brascamp-Lieb-Luttinger inequalities for convex domains of finite inradius, Duke Mathematical Journal, 113 (2002), 75, 93–131.
  • [19] P. Mörters, Y. Peres, Brownian Motion. Cambridge University Press, 2010.
  • [20] G. Pólya, G. Szegö, Isoperimetric Inequalities in Mathematical Physics. Annals of Mathematics Studies, no. 27, Princeton University Press, Princeton, N. J., 1951.
  • [21] S.C. Port, C.J. Stone,  Brownian Motion and Classical Potential Theory. Academic Press 1978.
  • [22] T. Ransford,  Potential Theory in the Complex Plane. Cambridge Univ. Press, 1995.
  • [23] G. Serafin,  Exit times densities of the Bessel process. Proc. Am. Math. Soc. 145(7) (2017), 3165-3178.
  • [24] A.S. Sznitman,   Brownian Motion, Obstacles and Random Media. Springer Science & Business Media, 1998.
  • [25] N. Zorii,  Precise estimate of the 2-capacity of a condenser. Ukrainian Math. J. 42(2) (1990), 224-228.