• Aucun résultat trouvé

Gradient estimates and Harnack inequalities on non-compact Riemannian manifolds

N/A
N/A
Protected

Academic year: 2021

Partager "Gradient estimates and Harnack inequalities on non-compact Riemannian manifolds"

Copied!
18
0
0

Texte intégral

(1)

www.elsevier.com/locate/spa

Gradient estimates and Harnack inequalities on non-compact Riemannian manifolds

I

Marc Arnaudon

b

, Anton Thalmaier

c

, Feng-Yu Wang

a,d,

aSchool of Mathematical Sciences, Beijing Normal University, Beijing 100875, China

bD´epartement de math´ematiques, Universit´e de Poitiers, T´el´eport 2 - BP 30179, F–86962 Futuroscope Chasseneuil Cedex, France

cInstitute of Mathematics, University of Luxembourg, 162A, avenue de la Fa¨ıencerie, L–1511 Luxembourg, Luxembourg dDepartment of Mathematics, Swansea University, Singleton Park, SA2 8PP, Swansea, UK

Received 1 October 2008; received in revised form 1 July 2009; accepted 1 July 2009 Available online 8 July 2009

Abstract

A gradient-entropy inequality is established for elliptic diffusion semigroups on arbitrary complete Riemannian manifolds. As applications, a global Harnack inequality with power and a heat kernel estimate are derived.

c

2009 Elsevier B.V. All rights reserved.

MSC:58J65; 58J35; 60H30

Keywords:Harnack inequality; Heat equation; Gradient estimate; Diffusion semigroup

1. The main result

LetM be a non-compact complete connected Riemannian manifold, andPt be the Dirichlet diffusion semigroup generated byL =∆+ ∇V for someC2functionV. We intend to establish reasonable gradient estimates and Harnack type inequalities forPt. In case that Ric−HessV is bounded below, a dimension-free Harnack inequality was established in [14] which, according

ISupported in part by WIMICS, NNSFC(10721091) and the 973-Project.

Corresponding author at: School of Mathematical Sciences, Beijing Normal University, Beijing 100875, China. Tel.:

+86 010 5880 8811.

E-mail addresses:wangfy@bnu.edu.cn,F.Y.Wang@swansea.ac.uk(F.-Y. Wang).

0304-4149/$ - see front matter c2009 Elsevier B.V. All rights reserved.

doi:10.1016/j.spa.2009.07.001

(2)

to [15], is indeed equivalent to the corresponding curvature condition. See e.g. [2] for equiva- lent statements on heat kernel functional inequalities; see also [8,3,7] for a parabolic Harnack inequality using the dimension–curvature condition by shifting time, which goes back to the classical local parabolic Harnack inequality of Moser [9].

Recently, some sharp gradient estimates have been derived in [11,18] for the Dirichlet semi- group on relatively compact domains. More precisely, forV =0 and a relatively compact open C2domainD, the Dirichlet heat semigroupPtDsatisfies

|∇PtDf|(x)≤C(x,t)PtDf(x), x∈ D, t>0, (1.1) for some locally bounded functionC:D×]0,∞[→]0,∞[and all f ∈Bb+, the space of bounded non-negative measurable functions onM. Obviously, this implies the Harnack inequality

PtDf(x)≤ ˜C(x,y,t)PtDf(y), t >0, x,y∈ D, f ∈Bb+, (1.2) for some functionC˜:M2×]0,∞[→]0,∞[. The purpose of this paper is to establish inequalities analogous to(1.1)and(1.2)globally on the whole manifoldM.

On the other hand however, both(1.1)and(1.2) are, in general, wrong for Pt in place of PtD. A simple counter-example is already the standard heat semigroup onRd. Hence, we turn to search for the following slightly weaker version of gradient estimate:

|∇Ptf(x)| ≤δ

Pt(f log f)−Pt flogPt f(x)+C(δ,x)

t∧1 Ptf(x),

x∈ M, t>0, δ >0, f ∈Bb+, (1.3)

for some positive functionC:]0,∞[×M →]0,∞[. When Ric −HessV is bounded below, this kind of gradient estimate follows from [2, Proposition 2.6] but is new without curvature conditions. In particular, it implies the Harnack inequality with power introduced in [14] (see Theorem 1.2).

Theorem 1.1. There exists a continuous positive function F on]0,1] ×M such that

|∇Ptf(x)| ≤δ (Pt flog f −Ptf logPtf) (x) +

F(δ∧1,x) 1 δ(t∧1)+1

+2δ

e

Pt f(x),

δ >0, x∈ M, t >0, f ∈Bb+. (1.4)

Theorem 1.2. There exists a positive function C∈C(]1,∞[×M2)such that (Ptf(x))α ≤(Pt fα(y))exp

2(α−1)

e +αC(α,x,y) αρ2(x,y)

(α−1)(t∧1)+ρ(x,y) , α >1, t>0, x,y∈ M, f ∈Bb+,

whereρ is the Riemannian distance on M. Consequently, for anyδ > 2there exists a positive function Cδ ∈ C([0,∞[×M)such that the transition density pt(x,y) of Pt with respect to µ(dx):=eV(x)dx , wheredx is the volume measure, satisfies

pt(x,y)≤ exp

−ρ(x,y)2/(2δt)+Cδ(t,x)+Cδ(t,y) qµ(B(x,√

2t))µ(B(y,√

2t)) , x,y∈M, t ∈]0,1[.

(3)

Remark 1.1. According to the Varadhan asymptotic formula for short time behavior, one has limt→04tlogpt(x,y)= −ρ(x,y)2, x6= y. Hence, the above heat kernel upper bound is sharp for short time, asδis allowed to approximate 2.

The paper is organized as follows: In Section2we provide a formula expressingPt in terms of PtDand the joint distribution of(τ,Xτ), whereXt is theL-diffusion process andτ its hitting time to∂D. Some necessary lemmas and technical results are collected. Proposition 2.5 is a refinement of a result in [18] to make the coefficient ofρ(x,y)/tsharp and explicit. In Section3 we use parallel coupling of diffusions together with Girsanov transformation to obtain a gradient estimate for Dirichlet heat semigroup. Finally, complete proofs of Theorems 1.1 and1.2 are presented in Section4.

To prove the indicated theorems, besides stochastic arguments, we make use of a local gradient estimate obtained in [11] forV =0. For the convenience of the reader, we include a brief proof for the case with drift in theAppendix.

2. Some preparations

Let Xs(x)be anL-diffusion process with starting pointxand explosion timeξ(x). For any bounded openC2domainD⊂Msuch thatx ∈ D, letτ(x)be the first hitting time of Xs(x)at the boundary∂D. We have

Pt f(x)=E

f(Xt(x))1{t<ξ(x)}, PtDf(x)=E

f(Xt(x))1{t<τ(x)}. LetptD(x,y)be the transition density ofPtD with respect toµ.

We first provide a formula for the density hx(t,z) of (τ(x),Xτ(x)(x)) with respect to dt⊗ν(dz), whereνis the measure on∂Dinduced byµ(dy):=eV(y)dy.

Lemma 2.1. Let K(z,x)be the Poisson kernel in D with respect toν. Then hx(t,z)=

Z

D

−∂tptD(x,y)

K(z,y) µ(dy). (2.1)

Consequently, the density s7→`x(s)of τ(x)satisfies the equation:

`x(s)= Z

D

−∂tptD(x,y)

µ(dy). (2.2)

Proof. Every bounded continuous function f:∂D→Rextends continuously to a functionhon D¯ which is harmonic inDand represented by

h(x)= Z

D

K(z,x)f(z) ν(dz).

Recall thatz7→ K(z,x)is the distribution density ofXτ(x)(x). Hence E[f(Xτ(x)(x))] =h(x)=

Z

D

K(z,x)f(z) ν(dz).

On the other hand, the identity h(x)=E[h(Xt∧τ(x)(x))]

(4)

yields h(x)=

Z

D

ptD(x,y)h(y) µ(dy)+ Z

D

ν(dz)Z t 0

hx(s,z)f(z)ds

= Z

D

ptD(x,y)Z

D

K(z,y)f(z)ν(dz)

µ(dy)+ Z

D

ν(dz)Z t 0

hx(s,z)f(z)ds

= Z

D

f(z)Z

D

ptD(x,y)K(z,y) µ(dy)+ Z t

0

hx(s,z)ds ν(dz),

which implies that K(z,x)=

Z

D

ptD(x,y)K(z,y) µ(dy)+ Z t

0

hx(s,z)ds. (2.3)

Differentiating with respect totgives hx(t,z)= −∂t

Z

D

ptD(x,y)K(z,y) µ(dy). (2.4)

Since∂tptD(x,y)is bounded on[ε, ε−1] × ¯D× ¯Dfor anyε ∈]0,1[, Eq.(2.1)follows by the dominated convergence.

Finally, Eq.(2.2)is obtained by integrating(2.1)with respect toν(dz). Lemma 2.2. The following formula holds:

Ptf(x)= PtDf(x)+ Z

]0,t]×D

Pt−sf(z)hx(s,z)dsν(dz)

= PtDf(x)+ Z

]0,t]×D

Pt−sf(z)PsD/2h.(s/2,z)(x)dsν(dz).

Proof. The first formula is standard due to the strong Markov property:

Ptf(x)=E

f(Xt(x))1{t<ξ(x)}

=E

f(Xt(x))1{t<τ(x)}

+E

f(Xt(x))1{τ(x)<t<ξ(x)}

= PtDf(x)+E E

f(Xt(x))1{τ(x)<t<ξ(x)}|(τ(x),Xτ(x)(x))

= PtDf(x)+ Z

]0,t]×D

Pt−sf(z)hx(s,z)dsν(dz). (2.5) Next, since

spsD(x,y)=L psD(·,y)(x)=L PsD/2psD/2(·,y)(x)

= Ps/D2(L ps/D2(·,y))(x)=Ps/D2(∂upuD(·,y)|u=s/2)(x), it follows from(2.1)that

hx(s,z)=PsD/2h.(s/2,z)(x). (2.6)

This completes the proof.

We remark that formula(2.6)can also be derived from the strong Markov property without invoking Eq.(2.1). Indeed, for anyu <sand any measurable set A ⊂∂D, the strong Markov

(5)

property implies that

Pτ(x) >s, Xτ(x)(x)∈ A =E

1{u<τ(x)}

Pτ(x) >s, Xτ(x)(x)∈ A|Fu

= Z

D

puD(x,y)P

τ(y) >s−u, Xτ(y)(y)∈ A µ(dy), and thus,

hx(s,z)=PuDh.(s−u,z)(x), s>u >0, x∈ D, z∈∂D.

Lemma 2.3. Let D be a relatively compact open domain andρ∂D be the Riemannian distance to the boundary∂D. Then there exists a constant C>0depending on D such that

P{τ(x)≤t} ≤Ceρ2D(x)/16t, x∈ D, t>0.

Proof. Forx ∈ D, let R := ρD(x)andρx the Riemannian distance function to x. Since D is relatively compact, there exists a constantc >0 such that Lρx2 ≤ cholds on Doutside the cut-locus ofx. Letγt := ρx(Xt(x)), t ≥ 0. By Itˆo’s formula, according to Kendall [6], there exists a one-dimensional Brownian motionbt such that

t2≤2

tdbt+cdt, t≤τ(x).

Thus, for fixedt >0 andδ >0, Zs :=exp

δ tγs2−δ

tcs−4δ2 t2

Z s 0

γu2du

, s≤τ(x) is a supermartingale. Therefore,

P{τ(x)≤t} = P

s∈[0,tmax]γs∧τ(x)≥ R

≤P

s∈[0,t]max Zs∧τ(x)≥eδR2/t−δc−4δ2R2/t

≤ exp

cδ−1

t(δR2−4δ2R2) . The proof is completed by takingδ:=1/8.

Lemma 2.4. On a measurable space(E,F,µ)˜ satisfyingµ(E) <˜ ∞, let f ∈ L1(µ)˜ be non- negative withµ(˜ f) > 0. Then for every measurable functionψ such thatψf ∈ L1(µ), there˜ holds:

Z

E

ψfdµ˜ ≤ Z

E

f log f

µ(˜ f)dµ˜+ ˜µ(f)log Z

E

eψdµ.˜ (2.7)

Proof. This is a direct consequence of [12] Lemma 6.45. We give a proof for completeness.

Multiplying f by a positive constant, we can assume thatµ(˜ f) = 1. IfR

Eeψdµ˜ = ∞, then (2.7)is clearly satisfied.

If R

Eeψdµ <˜ ∞, then since R

Eeψdµ˜ ≥ R

{f>0}eψdµ˜, we can assume that f > 0 everywhere. Now from the fact that eψ1f ∈L1(fµ)˜ , we can apply Jensen’s inequality to obtain

log Z

E

eψdµ˜

=log Z

E

eψ1 f fdµ˜

≥ Z

E

log

eψ1 f

fdµ˜

(6)

(note the right-hand-side belongs toR∪ {−∞}). To finish we remark that sinceψf ∈ L1(µ)˜ , Z

E

log

eψ1 f

fdµ˜ =

Z

E

ψf dµ˜ − Z

E

f log fdµ.˜

Finally, in order to obtain precise gradient estimate of the type(1.4), where the constant in front ofρ(x,y)/tis explicit and sharp, we establish the following revision of [18, Theorem 2.1].

Proposition 2.5.Let D be a relatively compact open C2domain in M and K a compact subset of D. For anyε >0, there exists a constant C(ε) >0such that

|∇logptD(·,y)(x)| ≤ C(ε)log(1+t−1)

√ t

+(1+ε)ρ(x,y)

2t ,

t∈]0,1[, x∈K, y∈ D. (2.8)

In addition, if D is convex, the above estimate holds forε=0and some constant C(0) >0.

Proof. Sinceδ:=minKρD >0, it suffices to deal with the case where 0<t ≤1∧δ. To this end, we combine the argument in [18] with relevant results from [16,17]. Lett∈(0,1∧δ],t0= t/2 andy∈ Dbe fixed, and take

f(x,s)=ps+tD

0(x,y), x∈ D, s>0.

(a) ApplyingTheorem A.1of theAppendixto the cube

Q:= B(x, ρD(x))× [s−ρD(x)2/2,s] ⊂D× [−t0,t0], s≤t0, we obtain

|∇log f(x,s)| ≤ c0

ρD(x)

1+log A f(x,s)

, s≤t0, (2.9)

whereA :=supQ f andc0>0 is a constant depending on the dimension and curvature onD.

By [7, Theorem 5.2], A≤c1f

x,s+ρD(x)2

, s∈]0,1], x∈ D, (2.10)

holds for some constantc1 > 0 depending on DandL. Moreover, by the boundary Harnack inequality of [4] (which treatsZ =0 but generalizes easily to non-zeroC1driftZ),

f

x,s+ρD(x)2

≤c2f(x,s), s∈]0,1], x∈D, (2.11) for some constantc2>0 depending onDandL. Combining(2.9)–(2.11), there exists a constant c>0 depending onDandLsuch that

|∇log f(x,s)| ≤ c

s, x∈ D, s∈]0,t0] withρD(x)2≤s. (2.12) (b) Let

Ω=n

(x,s):x∈ D, s∈ [0,t0], ρD(x)2≥so

andB=sup f. Since∂sf =L f, for any constantb≥1, we have (L−∂s)

f logb B f

= −|∇f|2 f .

(7)

Next, again by∂sf =L f and the Bochner–Weizenb¨ock formula, (L−∂s)|∇f|2

f ≥ −2k|∇f|2 f ,

wherek≥0 is such that Ric− ∇Z ≥ −konD. Then the function h:= s|∇f|2

(1+2ks)f − f logb B f satisfies

(L−∂s)h ≥0 onD×]0,∞[. (2.13)

Obviouslyh(·,0)≤0, and(2.12)yieldsh(x,s)≤0 fors=ρD(x)2provided the constantbis large enough. Then the maximum principle and inequality(2.13)implyh≤0 onΩ. Thus,

|∇log f(x,s)|2≤(2k+s−1)logb B

f , (x,s)∈Ω. (2.14)

(c) If Dis convex, by [16, Theorem 2.1] withδ = √

t and t = 2t0, we obtain (note the generator therein is 12L)

f(x,t0)= p2tD

0(x,y)= p2tD

0(y,x)≥c1ϕ(y)t0−d/2eρ(x,y)2/8t0, x∈K, y∈ D

for some constantc1 >0, whereϕ >0 is the first Dirichlet eigenfunction of L on D. On the other hand, the intrinsic ultracontractivity forPtDimplies (see e.g. [10])

f(z,s)=ps+tD

0(z,y)≤c2ϕ(y)t0(d+2)/2, z,y∈D, s≤t0,

for some constantc2>0 depending onD,K andL. Combining these estimates we obtain B

f(x,s) ≤c3t0−1eρ(x,y)2/8t0, x ∈K, s≤t0,

for some constantc3 > 0 depending onD, K andL. Hence by(2.14)for s = t0 we get the existence of a constantC >0 such that

|∇logp2tD

0(·,y)|2≤(t0−1+2k)

C+logt0−1+ρ(x,y)2 8t0

for ally∈ D,x∈ K andt0∈]0,1[witht0≤ρD(x)2. This completes the proof by noting that t =2t0.

(d) Finally, ifDis not convex, then there exists a constantσ >0 such that h∇XN,Xi ≥ −σ|X|2, X ∈T∂D,

whereN is the outward unit normal vector field of∂D, andT∂Dis the set of all vector fields tangent to∂D. Letψ ∈ C(D¯)such thatψ = 1 forρD ≥ ε, 1 ≤ ψ ≤ e2εσ forρD ≤ ε, andNlogψ|D ≥σ. By Lemma 2.1 in [17],∂Dis convex under the metricg˜:=ψ−2h·,·i. Let

∆˜,∇˜ andρ˜be respectively the Laplacian, the gradient and the Riemannian distance induced by g. By Lemma 2.2 in [17],˜

L:=∆+ ∇V =ψ−2h

∆˜ +(d−2)ψ∇ψi + ∇V.

(8)

SinceDis convex underg, as explained in the first paragraph in Section 2 of [17],˜ g˜(∇ ˜˜ρ(y,·),∇˜ϕ)|D <0,

so that

σ(y)˜ :=sup

D

g(˜ ∇ ˜˜ρ(y,·),∇˜ϕ) <∞, y∈ D.

Hence, repeating the proof of Theorem 2.1 in [16], but using ρ˜ and ∇˜ in place of ρ and∇ respectively, and taking into account thatψ→1 uniformly asε→0, we obtain

p2tD

0(x,y)≥ C1(ε)ϕ(y)t0−d/2e−C2(ε)ρ(˜x,y)2/8t0

≥ C1(ε)ϕ(y)t0−d/2e−C2(ε)C3(ε)ρ(x,y)2/8t0

for some constantsC1(ε),C2(ε),C3(ε) >1 withC2(ε),C3(ε)→1 asε→0. Hence the proof is completed.

3. Gradient estimate for Dirichlet heat semigroup using coupling of diffusion processes Proposition 3.1.Let D be a relatively compact C2domain in M. For every compact subset K of D, there exists a constant C=C(K,D) >0such that for allδ >0, t >0, x0 ∈ K and for all bounded positive functions f on M,

|∇PtDf(x0)| ≤δPtD

flog f

PtDf(x0)

(x0)+C 1

δ(t∧1)+1

PtDf(x0). (3.1) Proof. We assume thatt∈]0,1[, the other case will be treated at the very end of the proof.

We write∇V = Z so thatL = ∆+Z. Since PtD only depends on the Riemannian metric and the vector field Z on the domain D, by modifying the metric andZ outside of Dwe may assume that Ric− ∇Z is bounded below (see e.g. [13]); that is,

Ric− ∇Z ≥ −κ (3.2)

for some constantκ≥0.

Fixx0∈ K. Let f be a positive bounded function onMandXs a diffusion with generatorL, starting atx0. For fixedt≤1, let

v= ∇PtDf(x0)

|∇PtDf(x0)|

and denote byu 7→ϕ(u)the geodesics inMsatisfyingϕ(˙ 0)=v. Then d

du u=0

PtDf(ϕ(u))=

∇PtDf(x0) .

To formulate the coupling used in [1], we introduce some notations.

IfY is a semimartingale inM, we denote by dY its Itˆo differential and by dmY the martingale part of dY: in local coordinates,

dY =

dYi +1

ij k(Y)dhYj,Yki ∂

∂xi

(9)

whereΓij kare the Christoffel symbols of the Levi–Civita connection; if dYi =dMi+dAiwhere Miis a local martingale and Ai a finite variation process, then

dmY =dMi

∂xi.

Alternatively, ifQ(Y):TY0M →TY.Mis the parallel translation alongY, then dYt =Q(Y)td

Z .

0

Q(Y)−1s ◦dYs

t

and

dmYt =Q(Y)tdNt

whereNt is the martingale part of the Stratonovich integralRt

0 Q(Y)−1s ◦dYs. Forx,y∈M, andynot in the cut-locus ofx, let

I(x,y)=

d−1

X

i=1

Z ρ(x,y)

0

|∇e˙(x,y)Ji|2+

R(e˙(x,y),Ji)Ji+ ∇e˙(x,y)Z,e˙(x,y)

s ds (3.3) wheree˙(x,y)is the tangent vector of the unit speed minimal geodesice(x,y)and(Ji)di=1are Jacobi fields alonge(x,y)which together withe˙(x,y)constitute an orthonormal basis of the tangent space atxandy:

Ji(ρ(x,y))=Px,yJi(0), i=1, . . . ,d−1;

herePx,y:TxM →TyMis the parallel translation along the geodesice(x,y).

Letc∈]0,1[. Forh>0 but smaller than the injectivity radius of D, andt >0, letXhbe the semimartingale satisfyingX0h=ϕ(h)and

dXhs = PX

s,XshdmXs+Z(Xhs)ds+ξshds, (3.4) where

ξsh:=

h ct +κh

n(Xsh,Xs)

with n(Xsh,Xs) the derivative at time 0 of the unit speed geodesic from Xhs to Xs, and PX

s,Xsh:TXsM → TXh

sM the parallel transport along the minimal geodesic from Xs toXhs. By convention, we putn(x,x)=0 andPx,x =Id for allx∈ M.

By the second variational formula and(3.2)(cf. [1]), we have dρ(Xs,Xsh)≤

I(Xs,Xsh)− h ct −κh

ds≤ −h

ctds, s≤Th,

whereTh :=inf{s ≥ 0 : Xs = Xsh}. Thus,(Xs,Xsh)never reaches the cut-locus. In particular, Th≤ctand

Xs =Xsh, s≥ct. (3.5)

Moreover, we haveρ(Xs,Xhs)≤hand

sh|2≤h2

κ+ 1 ct

2

. (3.6)

(10)

We want to compensate the additional drift ofXhby a change of probability. To this end, let Msh = −

Z s∧ct 0

rh,PX

r,Xhr dmXrE , and

Rsh=exp

Msh−1 2[Mh]s

.

ClearlyRhis a martingale, and underQh=Rh·P, the processXhis a diffusion with generatorL. Lettingτ(x0)(resp.τh) be the hitting time of∂DbyX (resp. byXh), we have

1{t<τh}≤1{t<τ(x0)}+1{τ(x

0)≤th}. But, sinceXhs =Xs fors≥ct, we obtain

1{τ(x0)≤th}=1{τ(x0)≤ct}1{th}. Consequently,

1 h

PtDf(ϕ(h))−PtDf(x0)

= 1 h E

h

f(Xht)Rth1{t<τh}− f(Xt(0))1{t<τ(x0)}

i

≤ 1 h E

h

f(Xht)Rht1{t<τ(x0)}− f(Xt(0))1{t<τ(x0)}

i + 1

h E h

f(Xht)Rth1{τ(x0)ct}1{t<τh}

i, and sinceXht =Xtthis yields

1 h

PtDf(ϕ(h))−PtDf(x0)

≤E

f(Xt)1{t<τ(x0)}

1

h(Rth−1) +1

h E h

f(Xht)Rth1{τ(x0)≤ct}1{th}

i. (3.7)

The left hand side converges to the quantity to be evaluated ashgoes to 0. Hence, it is enough to find appropriate lim sup’s for the two terms of the right hand side. We begin with the first term.

Letting Ysh=

Msh−1 2[Mh]s

and noting thathn(Xrh,Xr),PX

r,XrhdmXri =

2 dbr up to the coupling timeThfor some one- dimensional Brownian motionbr, we have

Rth =exp

Mth−1 2[Mh]t

≤1+Mth−1

2[Mh]t+(Yth)2exp(Yth)

=1+Mth− Z t

0

sh|2ds+(Yth)2exp(Yth).

From the assumptions, exp(Yth)andYth/hhave all their moments bounded, uniformly inh >0.

Consequently, since f is bounded, lim sup

h→0

E

f(Xt)1{t<τ(x0)}

1 h

Z t 0

rh|2dr+(Yth)2exp(Yth)

=0, which implies

(11)

lim sup

h→0

E

f(Xt)1{t<τ(x0)}

1

h(Rth−1)

≤lim sup

h→0

E

f(Xt)1{t<τ(x0)}

1 h

Z s 0

rh,PXr,Xh r dmXr

E . UsingLemma 2.4and estimate(3.6), we have forδ >0

E

f(Xt)1{t<τ(x0)}

1 h

Z s 0

rh,PX

r,XrhdmXrE

≤δPtD

f log f

PtDf(x0)

(x0)

+δPtDf(x0)logE

1{t<τ(x0)}exp 1

δh Z ct

0

sh,PXs,Xh

sdmXsE

≤δPtD

f log f

PtDf(x0)

(x0)+δPtDf(x0)logE

exp 1

δ2h2 Z ct

0

ξsh

2

ds

≤δPtD

f log f

PtDf(x0)

(x0)+δPtDf(x0)ct δ2

1

c2t22

≤δPtD

f log f

PtDf(x0)

(x0)+ C0

cδtPtDf(x0),

whereC0 = 1+(cκ)2(recall thatt ≤ 1). Since the last expression is independent ofh, this proves that

lim sup

h→0

E

f(Xt)1{t<τ(x0)}

1

h(Rth−1)

≤δPtD

f log f

PtDf(x0)

(x0)+ C0

cδtPtDf(x0). (3.8)

We are now going to estimate lim sup of the second term in Eq.(3.7). By the strong Markov property, we have

E h

f(Xht)Rth1{τ(x0)≤ct}1{th}

i

=EQh

h

Pt−ctD f(Xhct)1{τ(x0)≤cth}

i

≤ kPt−ctD fkQh

nτ(x0)≤ct < τho

. (3.9)

Sinceρ(Xsh,Xs)≤hct−sct fors∈ [0,ct], we have on{τ(x0)≤ct< τh}: ρD(Xτ(hx

0))≤hct−τ(x0) ct . Fors∈ [0, τh−τ(x0)], define

Ys0=ρ(Xτ(hx0)+s, ∂D),

and for fixed smallε > 0 (butε >h), let S0 = inf{s ≥ 0, Ys0 = ε or Ys0 =0}. Since under Qhthe processXhs is generated byL, the drift ofρ(Xsh, ∂D)isLρ(·, ∂D)which is bounded in a neighborhood of∂D. Thus, for a sufficiently smallε >0, there exists aQh-Brownian motionβ started at 0, and a constantN >0 such that

Ys :=h ct−τ(x0)

ct +

s+N s≥Ys0, s∈ [0,S0]. Let

(12)

S =inf{u≥0, Yu=εorYu=0}. Taking into account that on{τ(x0)=u},

{YS00 =ε} ∪ {S0>ct−u} ⊂ {YS=ε} ∪ {S>ct−u}, we have foru ∈ [0,ct],

Qh n

ct< τh|τ(x0)=uo

≤Qh{YS0 =ε|τ(x0)=u} +Qh

S0≥ct−u|τ(x0)=u

≤Qh{YS=ε|τ(x0)=u} +Qh{S≥ct−u|τ(x0)=u}

≤Qh{YS=ε|τ(x0)=u} + 1

ct−uEQh[S|τ(x0)=u]. Now using the fact that e−N Ys is a martingale andYs2−2sa submartingale, we get

Qh{YS=ε|τ(x0)=u} = 1−e−N hct−uct 1−e−Nε ≤C1h and

EQh[S|τ(x0)=u] ≤ EQh

h

YS2|τ(x0)=u i

≤ ε2Qh{YS=ε|τ(x0)=u}

21−e−N hct−uct

1−e−Nε ≤C2h(ct−u) ct for some constantsC1,C2>0. Thus,

Qh n

ct< τh|τ(x0)=u o

≤C1h+ 1 ct−u C2

h(ct−u) ct

≤C1h+C3

h ct ≤C4

h t

for some constantsC3,C4>0 (recall thatt≤1). Denoting by`hthe density ofτ(x0)underQh, this implies

Qh

nτ(x0)≤ct< τho

= Z ct

0

`h(u)Qh{ct < τhh =u}du

≤ C4h t

Z ct 0

`h(u)du

=C4h

t Qh{τ(x0)≤ct}.

In terms of D−h = {x ∈ D, ρD(x) > h}andσh = inf{s > 0, Xhs ∈ ∂D−h}, we have σh≤τ(x0)a.s. Hence, byLemma 2.3,

Qh{τ(x0)≤ct} ≤Qh

h≤cto

≤Cexp

−ρD−h(ϕ(h)) 16ct

,

where we used thatXhs is generated byLunderQh. This implies Qh

nτ(x0)≤ct< τho

≤C5

h t exp

−ρD−h(ϕ(h)) 16ct

. (3.10)

Sinceh1 PtD(ϕ(h))−PtD(x0)

converges to|∇PtDf(x0)|, we obtain from(3.7)–(3.10),

(13)

|∇PtDf(x0)| ≤δPtD

f log f

PtDf(x0)

(x0)

+ C0

cδt PtDf(x0)+C5kPt−ctD fk1 t exp

−ρD(x0) 16ct

. (3.11)

Finally, as explained in steps (c) and (d) of the proof of Proposition 2.5, for any compact set K ⊂D, there exists a constantC(K,D) >0 such that

kPt−ctD fk≤eC(K,D)/tPtDf(x0), c∈ [0,1/2], x0∈ K, t ∈]0,1]. Combining this with(3.11), we arrive at

|∇PtDf(x0)| ≤δPtD

f log f

PtDf(x0)

(x0)+ C0

cδtPtDf(x0) +C5

1 t exp

−ρD(x0) 16ct

exp

C(K,D) t

PtDf(x0). (3.12) Finally, choosingcsuch that

0<c< 1

2 ∧dist(K, ∂D) 16C(K,D), we get for some constantC >0,

|∇PtDf(x0)| ≤δPtD

f log f

PtDf(x0)

(x0)+C 1

δt +1

PtDf(x0),

x0∈K, δ >0, (3.13)

which implies the desired inequality.

To finish we consider the case t > 1. From the semigroup property, we have PtDf = P1D(Pt−1D f). So lettingg=Pt−1D f and applying(3.13)togat time 1, we obtain

|∇PtDf(x0)| ≤δP1D glog g P1Dg(x0)

!!

(x0)+C 1

δ +1

P1Dg(x0).

Now usingP1Dg=PtDf, we get

|∇PtDf(x0)| ≤δP1D(glogg)(x0)−PtDf(x0)logPtDf(x0)+C 1

δ +1

PtDf(x0).

Lettingϕ(x)=xlogx, we have forz∈D glogg(z)=ϕ E

f(Xt−1(z))1{t−1<τ(z)}

≤ E

ϕ f(Xt−1(z))1{t−1<τ(z)}

=Eϕ(f)(Xt−1(z))1{t−1<τ(z)}

= Pt−1D (flog f)(z),

where we successively used the convexity ofϕand the fact thatϕ(0)=0. This implies

|∇PtDf(x0)| ≤δPtD

f log f

PtDf(x0)

(x0)+C 1

δ +1

PtDf(x0), which is the desired inequality fort>1.

(14)

4. Proof of Theorems1.1and1.2

Proof of Theorem 1.1. We assume that t ∈]0,1[ and refer to the end of the proof of Proposition 3.1for the caset >1. Fixingδ >0 andx0 ∈ M, we take R=160/(δ∧1). LetD be a relatively compact open domain withC2boundary containing B(x0,2R)and contained in B(x0,2R+ε)for some smallε > 0. By the countable compactness of M, it suffices to prove that there exists a constantC = C(D)such that (1.4)holds on B(x0,R)withC in place of F(δ∧1,x0). We now fixx ∈ B(x0,R),t ∈]0,1]and f ∈ Bb+. Without loss of generality, we may and will assume thatPtf(x)=1.

(a) LetPs(x,dy)be the transition kernel of theL-diffusion process, and forx ∈ D,z ∈ M, let

νs(x,dz)= Z

D

hx(s/2,y)Pt−s(y,dz) ν(dy),

whereνis the measure on∂Dinduced byµ(dy)=eV(y)dy. ByLemma 2.2we have Ptf(x)=PtDf(x)+

Z

]0,t]×D×M

psD/2(x,y)f(z)dsµ(dy)νs(y,dz).

Then

|∇Ptf(x)| ≤ |∇PtDf(x)| +

Z

]0,t]×D×M

|∇logpsD/2(·,y)(x)|psD/2(x,y)f(z)dsµ(dy)νs(y,dz)

=: I1+I2. (4.1)

(b) ByProposition 3.1, we have I1≤δPtD(flog f)(x)+δ

e+C 1

δt +1

, x∈ B(x0,R), t ∈]0,1[, δ >0 (4.2) for someC=C(D) >0.

(c) ByProposition 2.5withε=1, we have I2

Z

]0,t]×M×D

Clog(√e+s−1)

s +2ρ(x,y) s

psD/2(x,y)f(z)dsνs(y,dz)µ(dy) (4.3) for some C = C(D) > 0 and all t ∈]0,1]. Applying Lemma 2.4 to the measure µ˜ :=

psD/2(x,y)dsνs(y,dz)µ(dy)onE :=]0,t] ×M×Dso that µ(E˜ )=P(τ(x)≤t < ξ(x))≤1,

we obtain I2≤ δE

(flog f)(Xt(x))1{τ(x)≤t<ξ(x)}

e+δE

f(Xt(x))1{τ(x)≤t<ξ(x)}

×log Z

]0,t]×M×D

exp

Clog(e+s−1) δ√

s +2ρ(x,y) sδ

ds psD/2(x,y)νs(y,dz) µ(dy)

≤ δE(f log f)(Xt(x))1{τ(x)≤t<ξ(x)}

e+δE

f(Xt(x))1{τ(x)≤t<ξ(x)}

×log Z

]0,t]×M×D

exp A

δ +9R sδ

ds psD/2(x,y)νs(y,dz) µ(dy), (4.4)

(15)

where

A:=sup

r>0

C

rlog(e+r)−r <∞. We get

I2 ≤ δE

(f logf)(Xt(x))1{τ(x)≤t<ξ(x)}

+δ e +δE

f(Xt(x))1{τ(x)≤t<ξ(x)}

logE

exp(9R/δτ(x)) + A

δ

≤ δE

(f logf)(Xt(x))1{τ(x)≤t<ξ(x)}

e+δlogE

exp(9R/δτ(x)) +A

≤ δE

(f logf)(Xt(x))1{τ(x)≤t<ξ(x)}

+δlogE

"

exp

9R (δ∧1)τ(x)

δ∧1δ #

+A+δ e

=δE(f logf)(Xt(x))1{τ(x)≤t<ξ(x)} +(δ∧1)logE

exp

9R (δ∧1)τ(x)

+A+δ

e. (4.5)

ByLemma 2.3and noting thatρ(x)≥ R, we have E

exp

9R (δ∧1)τ(x)

≤1+E

9R

(δ∧1)τ(x)exp

9R (δ∧1)τ(x)

=1+ Z

0

9Rs (δ∧1)exp

9Rs (δ∧1)

d ds

−P{τ(x)≤s−1}

ds

=1+ 9R (δ∧1)

Z 0

9R (δ∧1)s+1

exp

9Rs (δ∧1)

P{τ(x)≤s−1}ds

≤1+ 9R (δ∧1)

Z 0

9R (δ∧1)s+1

exp

9Rs (δ∧1)

exp

−R2s 16

ds

=1+ 9R (δ∧1)

Z 0

9R (δ∧1)s+1

exp

−Rs (δ∧1)

ds

=1+9 Z

0

(9u+1)exp(−u)du =:A0, sinceR=160/(δ∧1). This along with(4.5)yields

I2≤δE

(f logf)(Xt(x))1{τ(x)≤t<ξ(x)}

+logA0+A+δ

e. (4.6)

The proof is completed by combining(4.6)with(4.1),(4.2)and(4.4).

Proof of Theorem 1.2. ByTheorem 1.1,

|∇Ptf(x)| ≤δ (Pt(f log f)(x)−(Ptf)(x)logPtf(x)) +

F(δ∧1,x) 1

δ(t∧1)+1

+2δ e

Ptf(x), δ >0, x∈M. (4.7)

(16)

Forα >1 andx 6=y, letβ(s)=1+s(α−1)and letγ:[0,1] → M be the minimal geodesic fromxtoy. Then| ˙γ| =ρ(x,y). Applying(4.7)withδ=αρ(x,y)α−1 , we obtain

d

dslog(Ptfβ(s))α/β(s)s)=α(α−1) β(s)2

Pt(fβ(s)log fβ(s))−(Ptfβ(s))logPtfβ(s) Ptfβ(s)s) + α

β(s)

h∇Ptfβ(s),γ˙si Ptfβ(s)s)

≥ αρ(x,y) β(s)Pt fβ(s)s)

α−1 αρ(x,y)

Pt(fβ(s)log fβ(s))−(Pt fβ(s))logPt fβ(s)s)

− |∇Ptfβ(s)s)|

≥ −F

α−1

αρ(x,y)∧1, γs

α2ρ2(x,y)

β(s)(α−1)(t∧1)+αρ(x,y) β(s)

−2(α−1) eβ(s)

≥ −C(α,x,y)

αρ2(x,y)

(α−1)(t∧1)+ρ(x,y)

−2(α−1) e where C(α,x,y) := sups∈[0,1]1αF

α−1

αρ(x,y)∧1, γs

. This implies the desired Harnack inequality.

Next, for fixedα∈]1,2[, let K(α,t,x)=supn

C(α,x,y):y∈B(x,√ 2t)o

, t >0, x ∈M.

NoteK(α,t,x)is finite and continuous in(α,t,x)∈]1,2[×]0,1[×M. Letp:=2/α. For fixed t∈]0,1[, the Harnack inequality gives fory∈B(x,√

2t), (Ptf(x))2≤(Ptfα(y))pexp

2(2−p)

e +2K(α,t,x) 2α α−1 +

√ 2t

. Then, choosingT >t such thatq :=p/2(p−1) <T/t,

µ B(x,√

2t) exp

−2(2−p)

e −2K(α,t,x) 2α α−1+

√ 2t

− t T −qt

(Ptf(x))2

≤ Z

B(x,

2t)(Ptfα(y))pexp

− ρ(x,y)2 2(T−qt)

µ(dy).

Similarly to the proof of [1, Corollary 3], we obtain that for anyδ >2, choosingα= 2δ

2+δ ∈]1,2[ such thatδ > 2−2α = p

p−1 >2, there is a constantc(δ) > 0 such that the following estimate holds:

Eδ(x,t):=

Z

M

pt(x,y)2exp

ρ(x,y)2 δt

µ(dy)

≤ expn

c(δ)K(α,t,x)(1+

√ 2t)o µ(B(x,√

2t)) , t >0, x∈ M.

By [5, Eq. (3.4)], this implies the desired heat kernel upper bound forCδ(t,x):=c(δ)K(α,t,x) (1+

2t).

(17)

Appendix

The aim of the Appendix is to explain that the arguments in Souplet–Zhang [11] and Zhang [18] for gradient estimates of solutions to heat equations work as well in the case with drift.

Theorem A.1. Let L =∆+Z for a C1vector field Z . Fix x0∈M and R, T, t0>0such that B(x0,R)⊂M. Assume that

Ric− ∇Z ≥ −K (A.1)

on B(x0,R). There exists a constant c depending only on d, the dimension of the manifold, such that for any positive solution u of

tu=Lu (A.2)

on QR,T :=B(x0,R)× [t0−T,t0], the estimate

|∇logu| ≤c 1

R +T−1/2+

√ K

1+log sup

QR,T

u u

holds on QR/2,T/2.

Proof. Without loss of generality, let N := supQT,Ru = 1; otherwise replaceu byu/N. Let f =loguandω= |∇f|2

(1−f)2. By(A.2)we have L f + |∇f|2−∂t f =0

so that

tω= 2h∇f,∇∂tfi

(1− f)2 +2|∇f|2tf (1− f)3

= 2h∇f,∇(L f + |∇f|2)i

(1− f)2 +2|∇f|2(L f + |∇f|2) (1− f)3

= 2h∇f,∇(∆f + |∇f|2)i

(1− f)2 +2|∇f|2(∆f + |∇f|2) (1− f)3 +2h∇fZ,∇fi +2 Hessf(∇f,Z)

(1− f)2 +2|∇f|2hZ,∇fi

(1− f)3 . (A.3)

Moreover,

Lω =∆ω+hZ,∇|f|2i

(1− f)2 +2|∇f|2hZ,∇fi (1− f)3

=∆ω+2 Hessf(∇f,Z)

(1− f)2 +2|∇f|2hZ,∇fi

(1− f)3 . (A.4)

Finally, by the proof of [11, (2.9)] with−kreplaced by Ric(∇f,∇f)/|∇f|2, we obtain

∆ω−

2h∇f,∇(∆f + |∇f|2)i

(1− f)2 +2|∇f|2(∆f + |∇f|2) (1− f)3

Références

Documents relatifs

In this paper, we study the power of Gaussian curvature flow of a compact convex hypersurface and establish its Harnack inequality when the power is negative.. In the Harnack

Using the coupling by parallel translation, along with Girsanov’s theorem, a new version of a dimension- free Harnack inequality is established for diffusion semigroups on

In particular, we want to emphasize that Stochastic Analysis provides nat- ural tools to derive local estimates in the sense that the gradient bound at given point depends only

A Harnack type inequality for certain complex

Keywords: Bounded mean oscillation; Doubling measure; Metric space; Nonlinear; p-harmonic; Poincaré inequality; Potential theory;.. Quasiminimizer;

The existence of the invariant bundles with exponential separation extends in a natural way a well known theorem on the existence of a spectral decomposition associated with

Annales de l’Institut Henri Poincaré - Analyse non linéaire.. 307 HARNACK INEQUALITIES FOR VARIATIONAL INTEGRALS. COROLLARY 1. 2) can be

Working exactly as in [7, Section 10], the forward in time Harnack inequality of Theorem 1.2 can be used to establish locally quantitative H¨older estimates for local, weak solutions