• No results found

Log-Harnack inequality for stochastic Burgers equations and applications

N/A
N/A
Protected

Academic year: 2021

Share "Log-Harnack inequality for stochastic Burgers equations and applications"

Copied!
15
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Log-Harnack inequality for stochastic Burgers equations and

applications

Citation for published version (APA):

Wang, F. Y., Wu, J. L., & Xu, L. (2010). Log-Harnack inequality for stochastic Burgers equations and applications. (Report Eurandom; Vol. 2010038). Eurandom.

Document status and date: Published: 01/01/2010

Document Version:

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website.

• The final author version and the galley proof are versions of the publication after peer review.

• The final published version features the final layout of the paper including the volume, issue and page numbers.

Link to publication

General rights

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain

• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement:

www.tue.nl/taverne

Take down policy

If you believe that this document breaches copyright please contact us at:

[email protected]

(2)

EURANDOM PREPRINT SERIES 2010-038

Log-Harnack Inequality for Stochastic Burgers Equations and Applications

Feng-Yu Wang, Jiang-Lun Wu, Lihu Xu ISSN 1389-2355

(3)

Log-Harnack Inequality for Stochastic

Burgers Equations and Applications

Feng-Yu Wang

a),b)

, Jiang-Lun Wu

b)

, Lihu Xu

c)

a) School of Math. Sci. and Lab. Math. Com. Sys., Beijing Normal University, Beijing 100875, China b) Department of Mathematics, Swansea University, Singleton Park, SA2 8PP, UK

c) PO Box 513, EURANDOM, 5600 MB Eindhoven, The Netherlands

September 28, 2010

Abstract

By proving an L2-gradient estimate for the corresponding Galerkin

approxima-tions, the log-Harnack inequality is established for the semigroup associated to a class of stochastic Burgers equations. As applications, we derive the strong Feller property of the semigroup, the irreducibility of the solution, the entropy-cost in-equality for the adjoint semigroup, and entropy upper bounds of the transition density.

AMS subject Classification: 60J75, 60J45.

Keywords: stochastic Burgers equation, log-Harnack inequality, strong Feller property, irreducibility, entropy-cost inequality, entropy bounds.

1

Introduction

Let T = R/(2πZ) be equipped with the usual Riemannian metric, and let dθ denote the Lebesgue measure on T. Then

H := ½ x ∈ L2(dθ) : kxk2 := Z T x(θ)dθ = 0 ¾

is a separable real Hilbert space with inner product

(4)

hx, yi :=

Z T

x(θ)y(θ)dθ.

For x ∈ C2(T), the Laplacian operator ∆ is given by ∆x = x00. Let (A, D(A)) be the

closure of (−∆, C2(T) ∩ H) in H, which is a positively definite self-adjoint operator on H. Then

V := D(A1/2), hx, yi

V := hA1/2x, A1/2yi

gives rise to a Hilbert space, which is densely and compactly embedded in H. By the integration by parts formula, for any x ∈ C2(T) we have

kxk2 V = Z T (xAx)(θ)dθ = Z T |x0(θ)|2dθ.

Moreover, for x, y ∈ C1(T) ∩ H, set B(x, y) := xy0. Then B extends to a unique bounded

bilinear operator B : V × V → H with (see Proposition 2.1 below)

(1.1) kBkV →H := sup

kxkV,kykV≤1

kB(x, y)k ≤√π.

Consider the following stochastic Burgers equation

(1.2) dXt= −

©

νAXt+ B(Xt)}dt + QdWt,

where ν > 0 is a constant, B(x) := B(x, x) for x ∈ V , Q is a Hilbert-Schmidt operator on H, and Wt is the cylindrical Brownian motion on H. According to to [4, Chapter 5]

(see also [6, Theorem 14.2.4]), for any x ∈ H, this equation has a unique solution with the initial X0 = x, which is a continuous Markov process on H and is denoted by Xtx from

now on. If moreover x ∈ V , then Xx

t is a continuous process on V (see Proposition 2.3

below). We are concerned with the associated Markov semigroup Pt given by

Ptf (x) := Ef (Xtx), x ∈ H, t ≥ 0

for f ∈ Bb(H), the set of all bounded measurable functions on H.

The purpose of this paper is to investigate regularity properties of Pt, such as strong

Feller property, heat kernel upper bounds, contractivity properties, and entropy-cost in-equalities. To do this, a powerful tool is the dimension-free Harnack inequality introduced in [14] for diffuions on Riemannian manifolds (see also [1, 2] for further development). In recent years, this inequality has been established and applied intensively in the study of SPDEs (see e.g. [10, 15, 9, 7, 8, 17] and references within). In general, this type of Harnack inequality can be formulated as

(5)

where α > 1 is a constant, Cα is a positive function on (0, ∞) × H2 with Cα(t, x, x) = 0,

which is determined by the underlying stochastic equation.

On the other hand, in some cases this kind of Harnack inequality is not available, so that the following weaker version (i.e. the log-Harnack inequality)

(1.4) Ptlog f (x) ≤ log Ptf (y) + Cα(t, x, y), f ≥ 1

becomes an alternative tool in the study. In general, according to [16, Section 2], (1.4) is the limit version of (1.3) as α → ∞. This inequality has been established in [11] and [16], respectively, for semi-linear SPDEs with multiplicative noise and the Neumann semigroup on non-convex manifolds.

As for the stochastic Burgers equation (1.2), by using A1+σ for σ > 1

2 to replace A (i.e. the hyperdissipative equation is concerned), the first and the third named authors established an explicit Harnack inequality of type (1.3) in [18], where a more general framework, which includes also the stochastic hyperdissipative Navier-Stokes equations, was considered. But, when σ ≤ 1

2, the known arguments (i.e. the coupling argument and gradient estimate) to prove (1.3) are no longer valid. Therefore, in this paper we turn to investigate the log-Harnack inequality for Ptassociated to (1.2), which also provides some

important regularity properties of the semigroup (see Corollary 1.2 below). Note that the stochastic Burgers equation does not satisfy the Lipschitz and monotone conditions required in [11], the present study can not be covered there.

To state our main result, we introduce the intrinsic norm induced by the diffusion part of the solution. For any x ∈ H, let

kxkQ := inf

©

kzkH : z ∈ H, Q∗z = x

ª

,

where Q∗ is the adjoint operator of Q, and we take kxk

Q = ∞ if the set in the right

hand side is empty. Moreover, let k · k and k · kHS denote the operator norm and the

Hilbert-Schmidt norm respectively for bounded linear operators on H.

Theorem 1.1. Assume that ν3 ≥ 4πkA−1/2Qk2. Then for any f ∈ B

b(H) with f ≥ 1,

(1.5) Ptlog f (x) ≤ log Ptf (y) +

2πkQk2 HSkx − yk2Q 1 − exp[−4π ν2kQk2HSt] exph4π ν2(kxk 2∨ kyk2)i

holds for t > 0 and x, y ∈ H.

Before introducing consequences of Theorem 1.1, let us recalled that the invariant probability measure of Ptexists, and any invariant probability measure µ satisfies µ(V ) =

1. These follows immediately since V is compactly embedded in H and due to the Itˆo formula one has

EkX0 tk2H + 2ν Z t 0 EkX0 sk2Vds ≤ kQk2HSt.

(6)

Next, for any two probability measures µ1, µ2 on H, let Wc1, µ2) be the transportation-cost between them with transportation-cost function

(x, y) 7→ c(x, y) := kx − yk2 Qexp h4π ν2(kxk 2∨ kyk2)i. That is, Wc(µ1, µ2) = inf µ∈C (µ12) Z H×H c(x, y)µ(dx, dy), where C (µ1, µ2) is the set of all couplings of µ1 and µ2. Finally, let

BV(x, r) =

©

z ∈ V : kz − xkV < r

ª

, x ∈ V, r > 0. Corollary 1.2. Assume that ν3 ≥ 4πkA−1/2Qk2.

(1) For any t > 0, Pt is intrinsic strong Feller, i.e.

lim

kx−ykQ→0

Ptf (y) = Ptf (x), x ∈ H, f ∈ Bb(H).

(2) Let µ be an invariant probability measure of Pt and let Pt∗ be the adjoint operator

of Pt w.r.t. µ. Then the entropy-cost inequality

µ((Pt∗f ) log Pt∗f ) ≤ 2πkQk 2 HS 1 − exp[−4π ν2kQk2HSt] Wc(f µ, µ), f ≥ 0, µ(f ) = 1

holds for all t > 0.

(3) Let k · kQ ≤ Ck · kV hold for some constant C > 0. Then

(1.6) P(Xty ∈ BV(x, r)) > 0, x, y ∈ V, t, r > 0.

Consequently, Pt has a unique invariant probability measure µ, which is fully

sup-ported on V , i.e. µ(V ) = 1 and µ(G) > 0 for any non-empty open set G ⊂ V . Furthermore, µ is strong mixing, i.e. for any f ∈ Bb(H),

lim

t→∞Ptf (x) = µ(f ), ∀ x ∈ V.

(4) Under the same assumption as in (3), Pt has a transition density pt(x, y) w.r.t. µ

on V such that the entropy inequalities

(1.7) Z V pt(x, z) log pt(x, z) pt(y, z) µ(dz) ≤ 2πkQk 2 HSc(x, y) 1 − exp[−4π ν2kQk2HSt] and (1.8) Z V

pt(x, y) log pt(x, y)µ(dy) ≤ − log

Z V exp · 2πkQk 2 HSc(x, y) 1 − exp[−4π ν2kQk2HSt] ¸ µ(dy) hold for all t > 0 and x, y ∈ V .

(7)

To prove the above results, we present in Section 2 some preparations including a brief proof of (1.1), a convergence theorem for the Galerkin approximation of (1.2), and the continuity of the solution in V . Finally, complete proofs of Theorem 1.1 and Corollary 1.2 are addressed in Section 3.

2

Some preparations

Obviously, (1.1) is equivalent to the following result.

Proposition 2.1. kB(x, y)k2 ≤ πkxk2

Vkyk2V holds for any x, y ∈ V .

Proof. Since RTx(θ)dθ = 0, there exists θ0 ∈ T such that x(θ0) = 0. For any θ ∈ T,

let γ : [0, d(θ0, θ)] → T be the minimal geodesic from θ0 to θ, where d(θ0, θ)(≤ π) is the Riemannian distance between these two points. By the Schwartz inequality we have

|x(θ)|2 = ¯ ¯ ¯ ¯ Z d(θ0,θ) 0 d dsx(γs)ds ¯ ¯ ¯ ¯ 2 ≤ d(θ0, θ) Z T |x0(ξ)|2dξ ≤ πkxk2 V. Therefore, kB(x, y)k2 = Z T |(xy0)(θ)|2dθ ≤ πkxk2 Vkyk2V.

Remark. From the proof we see that (1.1) is a property in one-dimension, since for

d ≥ 2 there is no any constant C ∈ (0, ∞) such that kxk2≤ C

Z Td

|∇x|2(θ)dθ, x ∈ C1(Td)

holds. The invalidity of (1.1) in high dimensions is the main reason why we only consider here the stochastic Burgers equation rather than the stochastic Navier-Stokes equation.

Next, due to that fact that to prove the log-Harnack inequality we have to apply the Itˆo formula for a reasonable class of reference functions which is, however, not available in infinite-dimensions, we need to make use of the finite-dimensional approximations. To introduce the Galerkin approximation of (1.2), let us formulate H by using the standard ONB {ek : k ∈ Z} for the complex Hilbert space L2(T → C; dθ), where

ek(θ) :=

1

e

ikθ, θ ∈ T.

Obviously, ∆ek= −k2ek holds for all k ∈ Z, and an element

x :=X

k∈Z

(8)

belongs to H if and only if x0 = 0, ¯xk= x−k for all k ∈ ˆZ := Z \ {0}, and P k∈ˆZ|xk|2 < ∞. Therefore, H =½ X k∈ˆZ xkek : ¯xk= x−k, X k∈ˆZ |xk|2 < ∞ ¾ .

For any m ∈ N, let

Hm =

©

x ∈ H : hx, eki = 0 for |k| > m

ª

,

which is a finite-dimensional Euclidean space. Let πm : H → Hm be the orthogonal

projection. Let Bm = πmB and Qm = πmQ. Consider the following stochastic differential

equation on Hm:

(2.1) dXt(m) = −©νAXt(m)+ Bm(Xt(m))

ª

dt + QmdWt.

Since coefficients in this equation are smooth and dkXt(m)k2 ≤ 2kQmk2HSdt + 2hX

(m)

t , QmdWti,

we conclude that for any x ∈ Hm this equation has a unique solution Xtm.x which is

non-explosive. Let

Pt(m)f (x) = Ef (Xtm,x), t ≥ 0, x ∈ Hm, f ∈ Bb(Hm).

In the spirit of [4, Theorem 5.7], the next result implies

(2.2) Ptf (x) = lim

m→∞P

(m)

t f (xm), x ∈ H, f ∈ Cb(H)

for {xm ∈ Hm}m≥1 such that xm → x in H.

Proposition 2.2. For any {xm ∈ Hm}m≥1 such that kx − xmkH → 0, we have kXtx−

Xm,xm

t k → 0 in probability as m → ∞. Consequently, (2.2) holds.

Proof. Simply denote Xt(m) = Xtm,xm and Xt= Xtx. It is easy to see that

(2.3) E

Z t 0

(kXsk2V + kXs(m)k2V)ds ≤ C(1 + t)

holds for some constant C > 0. By the Itˆo formula we have

kXt− Xt(m)k2 = −2 Z t 0 © νkXs− Xs(m)k2V + hB(Xs) − B(Xs(m)), Xs− Xs(m)i ª ds + ηt(m), (2.4)

(9)

where ηt(m) := kQ − Qmk2HSt + kx − xmk2+ 2 sup r∈[0,t] ¯ ¯ ¯ ¯ Z r 0 hXs− Xs(m), (Q − Qm)dWsi ¯ ¯ ¯ ¯, which goes to 0 as m → ∞. Since by (1.1)

|hB(x) − B(y), x − yi| = |hB(x, x − y) + B(x − y, y), x − yi| ≤ πkx − yk(kxkV + kykV)kx − ykV,

it follows from (2.4) that

kXr− Xr(m)k2 π ν Z r 0 kXs− Xs(m)k2(kXsk2V + kXs(m)k2V)ds + ηt(m), r ∈ [0, t]. Therefore, kXt− Xt(m)k2 ≤ ηt(m) exp · π ν Z t 0 (kXsk2V + kXs(m)k2V)ds ¸ .

Combining this with (2.3) we obtain that for any N > 0,kXt− Xt(m)k2 ≥ ηt(m)eN π/ν ¢ ≤ P µ Z t 0 (kXsk2V + kXs(m)k2V)ds ≥ N C(1 + t) N

which goes to 0 as N → ∞. Since ηt(m) → 0 as m → ∞, this implies that kXt−Xt(m)k →

0 in probability as m → ∞.

Finally, we have the following result for the continuity of the solution in V .

Proposition 2.3. For any x ∈ V , Xx

t is a continuous process in V .

Proof. For fixed x ∈ V and T > 0, we introduce the map

Y : C([0, T ]; V ) → C([0, T ]; V ),

such that for any u ∈ C([0, T ]; V ), {Yt(u)}t∈[0,T ] solves the deterministic equation

(2.5) Y˙t(u) = −

©

νAYt(u) + B(Yt(u) + ut)

ª

, Y0(u) = x.

Then Y (u) ∈ C([0, T ]; V ), see e.g. [13, Theorem 3.2] (the theorem is for 2D Navier-Stokes equation, so it is of course true for our case).

Next, let Zt= Z t 0 e−ν(t−s)AQdW s.

Since Q is Hilbert-Schmidt on H, Zt is a continuous process on V (see e.g. [5, Theorem

5.9]). Therefore, Xx

(10)

3

Proofs of Theorem 1.1 and Corollary 1.2

According to [11], the key step to prove the log-Harnack inequality for Pt(m)is the gradient of type

|QmDPt(m)f |2(x) ≤ (P

(m)

t |QmDf |2)(x)C(t, x), f ∈ Cb1(Hm)

for some continuous function C on (0, ∞) × Hm, where D is the gradient operator on Hm,

i.e. for any f ∈ C1(H

m), the element Df (x) ∈ Hm is determined by

hDf (x), hi = Dhf (x) := lim ε→0

f (x + εh) − f (x)

ε , h ∈ Hm.

To derive the desired gradient estimate, we need the following exponential estimate for

Xm

t .

Lemma 3.1. For any x ∈ Hm and t ≥ 0,

E exp · ν 2kA−1/2Q mk2 µ kXtm,xk2+ ν Z t 0 kXm,x s k2Vds ¶¸ ≤ exp · ν(kxk2+ kQ mk2HSt) 2kA−1/2Q mk2 ¸ .

Proof. By the Itˆo formula, we have

(3.1) dkXtm,xk2+ 2νkXm,x t k2Vdt = kQmkHS2 dt + 2hXtm,x, QmdWti. Let τn= inf © t ≥ 0 : kXtm,xk ≥ n}, n ∈ N. We have τn → ∞ as n → ∞. Let Mt(n)= Z t∧τn 0 hXm,x s , QmdWsi.

Then for any λ > 0

t 7→ exp£2λMt(n)− 2λ2hM(n)it

¤ is a martingale. Therefore, it follows from (3.1) that

E exp · λkXt∧τm,xnk2+ 2νλ Z t∧τn 0 kXsm,xk2Vds − 2λ2 Z t∧τn 0 kQ∗mXsm,xk2ds ¸ ≤ E exp · λ¡kxk2+ tkQmk2HS ¢ + 2λMt(n)− 2λ2hM(n)it ¸ = exp£λ(kxk2+ tkQ mk2HS) ¤ . (3.2) Noting that

(11)

kQ∗mxk = kQ∗mA−1/2A1/2xk ≤ kQ∗mA−1/2kkxkV = kA−1/2Qmk2kxkV, x ∈ Hm,

by letting n ↑ ∞ in (3.2) and taking

λ = ν

2kA−1/2Q mk2

,

we complete the proof.

Lemma 3.2. Let ν3 ≥ 4πkA−1/2Q

mk2. Then for any f ∈ Cb1(Hm),

kQmDPt(m)f k2(x) ≤ ¡ Pt(m)kQmDf k2 ¢ (x) exp · ν2 ¡ kxk2+ tkQmk2HS ¢¸

holds for all t ≥ 0 and x ∈ Hm.

Proof. Let h ∈ Hm. According to e.g. [4, Section 5.4],

DhXtm,x:= limε→0

Xtm,x+εh− Xtm,x

ε , t ≥ 0

exists and solves the ordinary differential equation d dtDhX m,x t = − © νADhXtm,x+ ˜Bm(Xtm,x, DhXtm,x) ª ,

where ˜Bm(x, y) := B(x, y) + B(y, x) for x, y ∈ Hm. By (1.1), this implies that

d dtkDhX m,x t k2V = −2νkADhXtm,xk2− 2hADhXtm,x, ˜Bm(Xtm,x, DhXtm,x)i 1 2νk ˜Bm(X m,x t , DhXtm,x)k2 ν kX m,x t k2VkDhXtm,xk2V. Therefore, kDhXtm,xk2V ≤ khk2V exp · ν Z t 0 kXm,x s k2Vds ¸ . Since ν3 ≥ 4πkA−1/2Q mk2 implies that ν2 2kA−1/2Qmk2 ν ,

by Lemma 3.1 and using the Jensen inequality we arrive at

(3.3) EkDhXtm,xk2V ≤ khk2V exp · ν ¡ kxk2+ tkQ mk2HS ¢¸ .

(12)

Consequently, by the dominated convergence theorem we obtain

(3.4) DhPt(m)f (x) = EhDf (Xtm,x), DhXtm,xi, f ∈ Cb1(Hm), x ∈ Hm, t ≥ 0.

On the other hand, we have

kQmDPt(m)f k2 = sup k˜hk≤1 hQmDPt(m)f, ˜hi2 = sup k˜hk≤1 hDPt(m)f, Q∗ m˜hi2 = sup khkQm≤1 |DhPt(m)f |2, (3.5) where khkQm := inf{kzk : z ∈ Hm, Q mz = h}

and khkQm = ∞ if the set on the right hand side is empty. Now, for any h ∈ Hm with khkQm ≤ 1, let {zn}n≥1⊂ H be such that Q

mzn= h and kznk ≤ 1 + n1. By (3.4) we have |DhPt(m)|2(x) = ¡ EhDf (Xtm,x), DhXtm,xi ¢2EhQmDf (Xtm,x), DznX m,x t i ¢2 ¡EkQmDf (Xtm,x)k2 ¢ EkDznX m,x t k2 = ¡ EkQmDf (Xtm,x)k2 ¢ EkA−1/2D znX m,x t k2V.

Combining this with (3.3) and (3.5) and letting n ↑ ∞, we complete the proof.

According to the L2-gradient estimate in Lemma 3.2, we are able to prove the log-Harnack inequality for Pt(m) as in [11].

Proposition 3.3. Let ν3 ≥ 4πkA−1/2Q

mk2. For any f ∈ Bb(Hm) with f ≥ 1,

Pt(m)log f (x) ≤ log Pt(m)f (y) + πkQmk

2 HSkx − yk2Qmexp[ ν2(kxk2∨ kyk2)] 1 − exp[−2π ν2kQmk2HSt] holds for all t > 0 and x, y ∈ Hm.

Proof. It suffices to prove for kx − ykQm < ∞. Let {zn} ⊂ Hm be such that Q∗mzn = x − y

and kznk2 ≤ kx − yk2Qm+ 1

n. Let γ ∈ C1([0, t]; R) such that γ(0) = 0, γ(t) = 1. Finally, let

xs = (x − y)γ(s) + y, s ∈ [0, t]. Then, by Lemma 3.2 we have (see [11, Proof of Theorem

(13)

Pt(m)log f (x) − log Pt(m)f (y) = Z t 0 d ds © P(m) s log P (m) t−sf ª (xs)ds = Z t 0 n 1 2P (m) s kQmD log Pt−s(m)f k2+ γ0(s)hx − y, DPs(m)log P (m) t−sf i o (xs)ds Z t 0 Ps(m) n 1 2kQmD log P (m) t−sf k2 + |γ0(s)| · kz nke2π(kxsk 2+kQ mk2HSs)/ν2kQ mD log Pt−s(m)f k o (xs)ds kznk2 2 Z t 0 |γ0(s)|2e4π(kxsk2+kQmk2HSs)/ν2ds. Since kxsk ≤ kxk ∨ kyk, by taking

γ(s) = 1 − exp[− ν2kQmk2HSs] 1 − exp[−4π ν2kQmk2HSt] , s ∈ [0, t] we obtain

Pt(m)log f (x) − log Pt(m)f (y) + 2πkQk

2 HSkznk2 1 − exp[−4π ν2kQk2HSt] exph4π ν2(kxk 2∨ kyk2)i.

This completes the proof by letting n → ∞.

Proof of Theorem 1.1. It suffices to prove for f ∈ Cb(H) with f ≥ 1. Let kx − ykQ < ∞.

For any ε > 0, let z ∈ H such that Q∗z = x − y and kzk2 ≤ kx − yk2

Q+ ε. For any m ∈ N,

we have Q∗

mz = πmx − πmy. Let xm = πmx, zm = πmz and ym = πmy + Q∗m(z − πmz).

Then zm ∈ Hm and Q∗mzm = xm− ym, so that

kxm− ymk2Qm ≤ kzmk

2 ≤ kx − yk2

Q+ ε.

Moreover, it is easy to see that xm → x and ym → y hold in H. Combining these with

Proposition 3.3 and (2.2), and letting m → ∞ and ε → 0, we complete the proof.

Proof of Corollary 1.2. The intrinsic strong Feller property follows from [16, Proposition 2.3], while the entropy-cost inequality in (2) follow from the proof of Corollary 1.2 in [11]. So, it remains to prove (3).

(a). Applying (1.5) to f := 1 + m1B(x,r) for m ≥ 1, we obtain

(3.6) Ptlog(1 + m1BV(x,r))(x) ≤ log ©

1 + mPt1BV(x,r)(y) ª

+ α(t)c(x, y), t > 0, m ≥ 1 for some function α : (0, ∞) → (0, ∞) independent of x, y and m. By Proposition 2.3 we have kXx

t − xkV → 0 as t → 0. Then there exists t0 > 0 depending only on x such that P(kXx

t − xkV < r) ≥

1

(14)

Thus, if P(Xty ∈ BV(x, r)) = 0 for some t ∈ (0, t0], then (3.6) yields that 1

2log(1 + m) ≤ Ptlog(1 + m1BV(x,r))(x) ≤ α(t)c(x, y), m ≥ 1,

which is impossible since k · kQ ≤ Ckx − ykV implies that c(x, y) < ∞ for x, y ∈ V .

Therefore,

P(Xtz ∈ BV(x, r)) > 0, t ∈ (0, t0], z ∈ V. Combining this with the Markov property we see that for t > t0,

P(Xty ∈ B(x, r)) =

Z

V

P(Xtz0 ∈ B(x, r))Pt−t0(y, dz) > 0, where Pt−t0(y, dz) is the distribution of X

y

t−t0. Therefore, (1.6) holds.

(b). The existence of the invariant measures for Burgers equation is well known ([6, Section 14.4]). Since (1) and k · kQ ≤ Ck · kV imply the strong Feller property of Pton V ,

by classical Doob’s Theorem, see Theorem 4.2.1 in [6], Eq. (1.2) has a unique invariant measure µ, which is strong mixing and supported on V . The full support property of µ, together with the strong Feller of Pt, implies the existence of transition density pt(x, y).

Finally, due to [16, Proposition 2.4(2)], (1.7) is equivalent to the log-Harnack inequality (1.5), while (1.8) follows from (1.5) according to the proof of [11, Corollary 1.2].

References

[1] M. Arnaudon, A. Thalmaier, F.-Y. Wang, Harnack inequality and heat kernel

es-timates on manifolds with curvature unbounded below, Bull. Sci. Math. 130 (2006),

223233.

[2] M. Arnaudon, A. Thalmaier, F.-Y. Wang, Gradient estimates and Harnack

inequal-ities on non-compact Riemannian manifolds, Stoch. Proc. Appl. 119 (2009), 3653–

3670.

[3] S. G. Bobkov, I. Gentil, M. Ledoux, Hypercontractivity of Hamilton-Jacobi equations, J. Math. Pures Appl. 80 (2001), 669–696.

[4] G. Da Prato, Kolmogorov equations for stochastic PDEs, Advanced Courses in Math-ematics. CRM Barcelona, Birkh¨auser Verlag, Basel, 2004.

[5] G. Da Prato and J. Zabczyk, Stochastic Equations in Infinite Dimensions, Cambridge Univ. Press, 1992.

[6] G. Da Prato and J. Zabczyk, Ergodicity for infinite-dimensional systems, London Mathematical Society Lecture Note Series, vol. 229, Cambridge University Press, Cambridge, 1996.

(15)

[7] G. Da Prato, M. R¨ockner, F.-Y. Wang, Singular stochastic equations on Hilbert

spaces: Harnack inequalities for their transition semigroups, J. Funct. Anal. 257

(2009), 992–017.

[8] A. Es-Sarhir, M.-K. v. Renesse and M. Scheutzow, Harnack inequality for functional

SDEs with bounded memory, Electron. Commun. Probab. 14 (2009), 560–565. [9] W. Liu, F.-Y. Wang, Harnack inequality and strong Feller property for stochastic

fast-diffusion equations, J. Math. Anal. Appl. 342 (2008), 651–662.

[10] M. R¨ockner, F.-Y. Wang, Harnack and functional inequalities for generalized Mehler

semigroups, J. Funct. Anal. 203 (2003), 237–261.

[11] M. R¨ockner, F.-Y. Wang, Log-harnack inequality for stochastic differential equations

in Hilbert spaces and its consequences, Infin. Dim. Anal. Quat. Probab. Relat. Top.

13 (2010), 27–37.

[12] F. Otto, C. Villani, Generalization of an inequality by Talagrand, and links with the

logarithmic Sobolev inequality, J. Funct. Anal. 173 (2000), 361–400.

[13] R. Temam, Navier-Stokes equations and nonlinear functional analysis, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, (1995), xiv+141. [14] F.-Y. Wang, Logarithmic Sobolev inequalities on noncompact Riemannian manifolds,

Probab. Theory Related Fields 109 (1997), 417–424.

[15] F.-Y. Wang, Harnack inequality and applications for stochastic generalized porous

media equations, Ann. Probab. 35 (2007), 1333–1350.

[16] F.-Y. Wang, Harnack inequalities on manifolds with boundary and applications, J. Math. Pures Appl. 94 (2010), 304–321.

[17] F.-Y. Wang, Harnack inequality for SDE with multiplicative noise and extension

to Neumann semigroup on non-convex manifolds, to appear in Annals of Probab.

arXiv:0911.1644

[18] F.-Y. Wang, L. Xu, Bismut type formula and its application to stochastic

Referenties

GERELATEERDE DOCUMENTEN

It tries to determine the sources of inequality for each of the three economic periods identified: the export-led growth period, the Import Substitution Industrialization period

Wanneer de Wasbeer zich permanent vestigt in Nederland zijn, gezien de ervaringen in Duitsland waar lokaal hoge dichtheden worden bereikt, de effecten met betrekking tot

Differentiaalvergelijkingen zijn een onderwerp dat zich goed laat behandelen met grafische methoden, waarin de differentiaalrekening echt wordt gebruikt, en dat zelf ook

The main purpose of this paper is to establish the Bismut type derivative formula for the Markov semigroup associated to stochastic Navier-Stokes type equations, and as applications,

the share in total income of the rich and the upper middle class declines and the share of the rest rises. 6.2.1

In this chapter, a brief introduction to stochastic differential equations (SDEs) will be given, after which the newly developed SDE based CR modulation model, used extensively in

Plant species Red List data were extracted from the following sources: (1) Global data: The IUCN Red List of Threatened Species (http://www.iucnredlist.org, accessed August 2009);

Naast de in artikel 2.10 bedoelde verpleging, omvat verpleging tevens de zorg zoals verpleegkundigen die plegen te bieden, zonder dat die zorg gepaard gaat met verblijf, en