• Tidak ada hasil yang ditemukan

getdoccd14. 118KB Jun 04 2011 12:04:59 AM

N/A
N/A
Protected

Academic year: 2017

Membagikan "getdoccd14. 118KB Jun 04 2011 12:04:59 AM"

Copied!
7
0
0

Teks penuh

(1)

ELECTRONIC

COMMUNICATIONS in PROBABILITY

A COUNTEREXAMPLE FOR THE OPTIMALITY OF

KENDALL-CRANSTON COUPLING

KAZUMASA KUWADA

Department of Mathematics, Faculty of Science, Ochanomizu University, Tokyo 112-8610, Japan

email: kuwada@math.ocha.ac.jp

KARL-THEODOR STURM

Institute for Applied Mathematics, University of Bonn, Wegelerstrasse 6, 53115 Bonn, Ger-many

email: sturm@uni-bonn.de

Submitted November 24 2005, accepted in final form September 29 2006 AMS 2000 Subject classification: 60D05, 58J65

Keywords: Brownian motion, manifold, optimal coupling, Kendall-Cranston coupling

Abstract

We construct a Riemannian manifold where the Kendall-Cranston coupling of two Brownian particles does not maximize the coupling probability.

1

Introduction

Given two stochastic processesXtandYton a state space M, a couplingZt= (Zt(1), Z

(2)

t ) is

a process on M ×M so that Z(1) or Z(2) has the same distribution as X or Y respectively. Of particular interest in many applications is the distribution of the coupling time T(Z) := inf{t >0 ;Zs(1)=Zs(2) for alls > t}. The goal is to make the coupling probabilityP[T(Z)≤t]

as large as possible by taking a suitable coupling. When X and Y are Brownian motions on a Riemannian manifold, Kendall [3] and Cranston [1] constructed a coupling by using the Riemannian geometry of the underlying space. Roughly speaking, under their coupling, infinitesimal motion ∆Yt∈TYtMat timetis given as a sort of reflection of ∆Xtvia the minimal

geodesic joining Xt and Yt. Their coupling has the advantage of controlling the coupling

probability by using geometric quantities such as the Ricci curvature. As a result, Kendall-Cranston coupling produces various estimates for heat kernels, harmonic maps, eigenvalues etc. under natural geometric assumptions.

On the other hand, there is the question of optimality. We say that a couplingZ ofX andY

isoptimal at time tif

P[T(Z)≤t]≥P[T( ˜Z)≤t]

holds for any other coupling ˜Z. Though Kendall-Cranston coupling has a good feature as mentioned, in general there is no reason why it should be optimal.

(2)

The Kendall-Cranston coupling is optimal if the underlying space has a good symmetry. For example, in the caseM =Rd, the Kendall-Cranston coupling (Z(1), Z(2)) is nothing but the mirror coupling. It means that Zt(2) = Ψ(Z

(1)

t ) up to the time they meet, where Ψ is a

reflection with respect to a hyperplane inRd so that Ψ(X0) =Y0. It is well known that the

mirror coupling is optimal. Indeed, it is the only coupling which is optimal and Markovian [2]. More generally, the same result holds if there is a sort of reflection structure like a map Ψ on Rd (see [4]).

In this paper, we show that the Kendall-Cranston coupling isnot optimal in general.

Theorem 1.1 For eacht >0, there is a complete Riemannian manifoldM where the Kendall-Cranston coupling of two Brownian motions X· and Y· with specified starting points is not

optimal.

The proof of Theorem 1.1 is reduced to the case t = 1 by taking a scaling of Riemannian metric. We construct a manifoldM in the next section and prove Theorem 1.1 in section 3.

Notation: Given a Riemannian manifold N we denote by BN

r (x) or simply Br(x) the open

ball inN of radiusrcentered atx.

Given a Brownian motion (Xt)t≥0 onN we denote byτA= inf{t >0 : Xt∈A} the hitting

time of a setA⊂N. We remark that, throughout this article,τAalways stands for the hitting

time for the process (Xt)t≥0even when we consider a coupled motion (Xt, Yt)t≥0.

2

Construction of the manifold

We take three parameter R > 0, ζ > 0 and δ > 0 such that ζ < R/4 and δ < ζ/3. Let

C=R×S1 be a cylinder with a flat metric such that the length of a circleS1equals ζ. For simplicity of notation, we write z = (r, θ) for z ∈ C where r ∈ R and θ ∈ (−ζ/2, ζ/2] such that the Riemannian metric is written asdr2+2. If appropriate, anyθRwill be regarded modζand considered as element of (−ζ/2, ζ/2]. We put

M1:= ([−R,∞)×S1)\BCδ((0, ζ/2))⊂C

and write∂1,0:=∂BδC((0, ζ/2)) as well as∂1,2:={−R} ×S1 (see Fig.1). LetC′ be a copy of

C. Then we put analogously

M2:= ((−∞, R]×S1)\BC ′

δ ((0,0))⊂C′

and write∂2,0:=∂BC ′

δ ((0,0)) as well as∂2,1:={R} ×S1. LetM0=S1×[−1,1] be another cylinder. We write z∈M0 byz= (ϕ, r) whereϕ∈(0,2π] andr∈[−1,1]. Now we define a

C∞-manifoldM (see Fig.2) by M =M

0⊔M1⊔M2/∼, where the identification “∼” means

∂1,2∋(−R, θ)∼(R, ζ/2−θ)∈∂2,1 forθ∈(−ζ/2, ζ/2],

∂1,0∋(δcosϕ, ζ/2−δsinϕ)∼(ϕ,−1)∈M0 forϕ∈(0,2π],

∂2,0∋(δcosϕ, δsinϕ)∼(ϕ,1)∈M0 forϕ∈(0,2π].

We endowM with aC∞-metricgsuch that (M, g) becomes a complete Riemannian manifold

and:

(3)

Fig.1

≈ ≈

b

M

1

R

∂1,0

∂1,2

ζ

(0,0)

C

Fig.2

M

M0

M1

M2

b b

x

y ∂2,1=∂1,2

b

b Φ

≈ ≈

(ii) g|M2 coincides with the metric onM2 inherited fromC ′,

(iii) g|M0 is invariant under maps (θ, r)7→(θ,−r) and (θ, r)7→(θ+ϕ, r) onM0,

(iv) d((−1,0),(1,0)) =ζforz1= (−1,0), z2= (1,0)∈M0

where dis the distance function onM.

3

Comparison of coupling probabilities

LetM be the manifold constructed above (with suitably chosen parametersR,ζ and δ) and fix two points x= (0, ζ/6)∈M1 andy= (0, ζ/3)∈M2.

In this paper, the construction of Kendall-Cranston coupling is due to von Renesse [5]. We will try to explain his idea briefly. His approach is based on the approximation by coupled geodesic random walks {Ξˆk}

k∈Nstarting in (x, y) whose sample paths are piecewise geodesic.

Given their positions after (n−1)-th step, one determines its next directionξn according to

the uniform distribution on a small sphere in the tangent space and the other does it as the reflection ofξn along a minimal geodesic joining their present positons. We obtain a

Kendall-Cranston coupling (Xt, Yt) by taking the (subsequential) limit in distribution of them. We

will construct another Brownian motion ( ˆYt)t≥0onM starting iny, again defined on the same probability space as we construct (Xt, Yt) such that

P(X andY meet before time 1) < PX and ˆY meet before time 1.

In other words, ifQdenotes the distribution of (X, Y) and ˆQdenotes the distribution of (X,Yˆ) then

Proposition 3.1 Q[T ≤1]<Qˆ[T ≤1].

(4)

(i) ˆYt= Φ(Xt) fort∈[0, τ∂1,0∧T);

(ii) X and ˆY move independently fort∈[τ∂1,0, T) in caseτ∂1,0 < T;

(iii) ˆYt=Xtfort∈[T,∞).

Note thatτ∂1,2 =T holds whenτ∂1,2 ≤τ∂1,0 under ˆQ.

SetH =S1× {0} ⊂M

0⊂M. Forz1, z2∈M andA⊂M, minimal length of paths joiningz1 andz2 which intersectAis denoted byd(z1, z2;A). We define a constantL0 by

L0:= inf

L∈(δ, R] ; d(z1, z2;H)≥d(z1, z2;∂1,2)

for somez1= (L, θ)∈M1, z2= (L, ζ/2−θ)∈M2

.

Lemma 3.2 R−ζ < L0< R.

Proof. First we show L0 < R. Let z1 = (R,0) ∈M1 and z2 = (R, ζ/2)∈ M2. Obviously there is a path of length 2Rjoining z1andz2 across∂1,2. Thus we haved(z1, z2;∂1,2)≤2R.

By symmetry ofM,

d(z1, z2;H) = 2d(z1, H) = 2

d(z1, ∂1,0) +

ζ

2

= 2pR2+ζ2/4δ+ζ >2R,

where the second equality follows from the third and fourth properties of g and the last inequality follows from the choice ofδ. These estimates implyL0< R.

Next, letz′

1= (R−ζ, θ)∈M1andz2′ = (R−ζ, ζ/2−θ)∈M2. In the same way as observed above, we have

d(z1′, z′2;H) = 2

p

(R−ζ)2+θ2δ+ζ2R2δ.

Note that the length of a path joiningz′

1andz′2which intersects both of∂1,2andHis obviously greater than d(z′

1, z2′ ;H). Thus, in estimatingd(z1′, z2′ ;∂1,2), it is sufficient to consider all

paths joiningz′

1 andz2′ across∂1,2which do not intersectH. Such a path must intersect both

{δ} ×S1M

1and{−δ} ×S1⊂M1(see Fig.3). Thus we have

d(z′1, z2′ ;∂1,2)≥d(z1′,{δ} ×S1) +d({−δ} ×S1, ∂1,2) +d(∂2,1, z2′)

≥(R−ζ−δ) + (R−δ) +ζ

= 2R−2δ.

Hence, the conclusion follows.

Set M′

1 := M1∩[−L0, L0]×S1 ⊂ C and M2′ := M2∩[−L0, L0]×S1 ⊂ C′. We define a submanifoldM′M with boundary byM=M

0⊔M1′⊔M2′/∼(see Fig.4). Let Ψ : M′→M′ be the reflection with respect toH. For instance, forz= (r, θ)∈M′

1, Ψ(z) = (r, ζ/2−θ)∈M2′. Note that Ψ is an isometry, Ψ◦Ψ = id and {z∈M′; Ψ(z) =z}=H.

LetX′be the given Brownian motion starting inxand now stopped at∂M, i.e. X

t=Xt∧τ∂M′. Define a stopped Brownian motion starting iny byY′

t = Ψ(Xt′) for t < τH and byYt=Xt

for t ≥τH (that is, the two Brownian particles coalesce after τH). Then we can prove the

(5)

Fig.3

Proof. Note that the minimal geodesic inM joiningz and Ψ(z) must intersectH for every

z∈M′by virtue of the choice ofL0. Thus, by the symmetry ofMwith respect toH, coupled

geodesic random walks ˆΞk are inE defined by

E:=n(z(1)· , z·(2))∈C([0,∞)→M ×M) ; z(2)t = Ψ(z

surely by taking a (subsequential) limit in distribution of{Ξˆk}

k∈N. Thus the conclusion follows.

The remark after the definition of ˆQimplies

ˆ

(6)

Next we give an upper estimate ofQ[T ≤1]. Let E:=

c under Q. In order to collide two Brownian motions

starting atXτ∂M′ and Ψ(Xτ∂M′), either of them must escape from the flat cylinder of length 2(L0−δ) where its starting point has distance L0−δfrom the boundary. This observation together with the strong Markov property yields

Q[{τ∂M′ < T ≤1} ∩Ec] =Q

Here the last inequality follows from Lemma 3.2. Consequently, we obtain

Q[T ≤1]≤PR2

≈0. Then Proposition 3.1 follows from (3.1) and (3.2).

Acknowledgment. This work is based on the discussion when the first-named author stayed in University of Bonn with the financial support from the Collaborative Research Center SFB 611. He would like to thank the institute for hospitality.

References

[1] Cranston, M.,“Gradient estimates on manifolds using coupling”, J. Funct. Anal.99(1991), no.1, 110–124. MR1120916

(7)

[3] Kendall, W.,“Nonnegative Ricci curvature and the Brownian coupling property”, Stochas-tics19(1986), 111–129 MR0864339

[4] Kuwada, K.,“On uniqueness of maximal coupling for diffusion processes with a reflection”, to appear in Journal of Theoretical Probability

Referensi

Dokumen terkait

The last term in the horizontal part of the reduced geodesic spray is related to the curvature of the horizontal distribution, at least in the case when M is a Riemannian manifold

Reflecting Brownian motion can be constructed as a strong Markov process not only on a domain with smooth or Lipschitz boundary with H¨ older cusps (see Lions and Sznitman [10],

(2007) use ingenious arguments based on ideas from differential games to show that a bounded convex planar domain cannot support any shy couplings of reflected Brownian motions if

We first show that Brownian motions conditioned not to intersect have a good chance of being reasonably far apart, and conclude that the corresponding simple walks are also far

Indeed, it seems that case (a) should give us the right picture: in the limit, in between large events lineages perform Brownian motions on the torus of sidelength 1 due to

the time reversed process is a drifted fractional Brownian motion, which continuously extends the one obtained in the theory of time reversal of Brownian diffusions when H = 1/2..

Keywords Wiener space, Sobolev space, Stochastic integral, Fractional Brownian Motion, Itˆ o’s formula.. AMS Subject Classification (1991)

Keywords: Malliavin calculus, Clark-Ocone formula, Brownian local time, Knight theorem, central limit theorem, Tanaka