• Aucun résultat trouvé

A sharp lower bound for the first eigenvalue on Finsler manifolds

N/A
N/A
Protected

Academic year: 2022

Partager "A sharp lower bound for the first eigenvalue on Finsler manifolds"

Copied!
14
0
0

Texte intégral

(1)

Ann. I. H. Poincaré – AN 30 (2013) 983–996

www.elsevier.com/locate/anihpc

A sharp lower bound for the first eigenvalue on Finsler manifolds

Guofang Wang

a

, Chao Xia

a,b,

aAlbert-Ludwigs-Universität Freiburg, Mathematisches Institut, Eckerstr. 1, 79104 Freiburg, Germany bMax-Planck-Institut für Mathematik in den Naturwissenschaften, Inselstr. 22, 04103, Leipzig, Germany

Received 22 November 2011; accepted 7 December 2012 Available online 11 January 2013

This paper is dedicated to Professor S.S. Chern on the occasion of his 100th birthday

Abstract

In this paper, we give a sharp lower bound for the first (nonzero) Neumann eigenvalue of Finsler-Laplacian in Finsler manifolds in terms of diameter, dimension, weighted Ricci curvature.

©2013 Elsevier Masson SAS. All rights reserved.

MSC:35P15; 53C60; 35A23

Keywords:Finsler-Laplacian; First eigenvalue; Weighted Ricci curvature; Poincaré inequality

1. Introduction

The study of the first (nonzero) eigenvalue of Laplacian in Riemannian manifolds plays an important role in differ- ential geometry. The first result on this subject, due to Lichnerowicz [11], says that for an n-dimensional smooth compact manifold without boundary, the first eigenvalue λ1 can be estimated below by nn1K, provided that its RicK >0. In this case, Obata[13] established a rigidity result, asserting the optimality of Lichnerowicz’ esti- mate. Namely, λ1=K if and only ifM is isometric to then-dimensional sphere with constant curvature n11K.

WhenK=0, Li–Yau[9,10]developed a method, which depends on the gradient estimate of the eigenfunctions, to give the lower bound of the first eigenvalue via diameterd, precisely,λ12dπ22. Their method had been improved by Zhong–Yang[24]to obtainλ1πd22, which is optimal in the sense that equality can be attained for one-dimensional circle. Very recently, Hang and Wang showed thatλ1>π2

d2 in[7], if the dimensionn >1. These results also hold true whenMis a manifold with convex boundary. WhenMis a convex domain inRn, this is a classical result of Payne–

Weinberger[17]. Later Chen–Wang[5]and Bakry–Qian[3]combined these results into a same framework, and gave estimates for the first eigenvalue of very general elliptic symmetric operators, via diameter and Ricci curvature. This sharp estimate on Riemannian manifolds has been also generalized to Alexandrov spaces by Qian–Zhang–Zhu[18].

GW is partly supported by SFB/TR71 “Geometric partial differential equations” of DFG. CX is supported by China Scholarship Council.

* Corresponding author.

E-mail addresses:guofang.wang@math.uni-freiburg.de(G. Wang),chao.xia@mis.mpg.de(C. Xia).

0294-1449/$ – see front matter ©2013 Elsevier Masson SAS. All rights reserved.

http://dx.doi.org/10.1016/j.anihpc.2012.12.008

(2)

Finsler geometry attracts many attentions in recent years, since it has broader applications in nature science. Simul- taneously Finsler manifold is one of the most natural metric measure spaces, which plays an important role in many aspects in mathematics. There exists a natural Laplacian on Finsler manifolds, which we call hereFinsler-Laplacian.

Unlike the usual Laplacian, the Finsler-Laplacian is a nonlinear operator. The objective of this paper is to study the lower bound for the first (nonzero) eigenvalue of this Finsler-Laplacian on Finsler manifolds. In[14]Ohta introduced the weighted Ricci curvature RicN for N ∈ [n,∞] of Finsler manifolds, following the work of Lott–Villani [12]

and Sturm[20]on metric measure space. He proved the equivalence of the lower boundedness of theRicN and the curvature-dimension conditionsCD(K, N )in[12,20]. As a byproduct, he obtained a Lichnerowicz type estimate on the first eigenvalue of Finsler-Laplacian under the assumptionRicNK >0. Another interesting type of eigenvalue estimates was obtained by Ge–Shen in[6], namely the Faber–Krahn type inequality for the first Dirichlet eigenvalue of the Finsler-Laplacian holds. See also[4]and[22]. Recently, we proved in[21]the Li–Yau–Zhong–Yang type sharp estimate for a so-called anisotropic Laplacian on a Minkowski space, which could be viewed as the simplest, but interesting and important case of non-Riemannian Finsler manifolds. In this paper, we shall generalize the results in [21]to general Finsler manifolds. Moreover, as in [5]and[3], we shall put the Li–Yau–Zhong–Yang type and the Lichnerowicz type sharp estimates into a uniform framework.

Our main result of this paper is

Theorem 1.1.Let(Mn, F, m)be ann-dimensional compact Finsler measure space, equipped with a Finsler structure F and a smooth measure m, without boundary or with a convex boundary. Assume that RicN K for some real numbersN ∈ [n,+∞]and K∈R. Letλ1be the first(nonzero)Neumann eigenvalue of the Finsler-Laplacianm, i.e.,

mu=λ1u, inM, (1)

with a Neumann boundary condition

u(x)Tx(∂M), (2)

if∂Mis not empty. Then

λ1λ1(K, N, d), (3)

whered is the diameter ofM,λ1(K, N, d)represents the first(nonzero)eigenvalue of the1-dimensional problem vT (t )v= −λ1(K, N, d)v in

d 2,d

2

, v

d 2

=v

d 2

=0, (4)

withT (t )varying according to different values ofKandN.T is explicitly defined by

T (t )=

⎧⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

(N−1)Ktan(

K

N1t ), forK >0, 1< N <,

−√

(N−1)Ktanh(

NK1t ), forK <0, 1< N <,

0, forK=0, 1< N <,

Kt, forN= ∞.

(5)

The precise definition of the Finsler measure space, convex boundary, diameterd, weighted Ricci curvatureRicN, gradient vector field∇, Finsler-Laplacianmwill be given in Section 2 below.

Equivalently, Theorem1.1gives an optimal Poincaré inequality in Finsler manifolds.

Theorem 1.2.Under the same assumptions as in Theorem1.1, we have

M

F2(u) dmλ1(K, N, d)

M

(u− ¯u)2dm, (6)

whereu¯is the average ofu.

In the case ofK >0 andN=n, Theorem1.1sharpens the Lichnerowicz type estimate given by Ohta[14], since by Meyer theorem, d (nπ1)K, and λ1(K, N, d)λ1(K, n, π

(n1)K)=nnK1. In the case ofK=0 andN=n,

(3)

Theorem1.1gives the Li–Yau–Zhong–Yang type sharp estimate for the Finsler-Laplacian, sinceλ1(0, n, d)=πd22. We remark that the Minkowski space(Rn, F )equipped with then-dimensional Lebesgue measure satisfies thatRicNK withK=0 andN =n(see e.g.[23]theorem on page 908 and[14, Theorem 1.2]), hence Theorem1.1covers the estimate in[21].

Our proof goes along the line of Bakry–Qian[3]. The technique is based on a comparison theorem on the gradient of the first eigenfunction with that of a one-dimensional (1-D) model function (Theorem3.1), which was developed by Kröger[8]and improved by Chen–Wang[5]and Bakry–Qian[3]. By using a Bochner–Weitzenböck formula es- tablished recently by Ohta–Sturm[16], we find that the one-dimensional model coincides with that in the Riemannian case, as presented in Theorem1.1. It should be not so surprising, because when we considerF inR, it can only be two pieces of linear functions. Since the 1-D model has been extensively studied in[3], it also eases our situation, although we deal with a nonlinear operator. One difficulty arises when we deal with the Neumann boundary prob- lem, since the convexity of boundary could not be directly applied due to the difference between the metric induced from the boundary itself and the metric induced from the gradient of the first eigenfunction. We will establish some equivalence between them (see Lemmas3.1and3.2) to overcome this difficulty. Another ingredient is a comparison theorem on the maxima of eigenfunction with that of the 1-D model function (Theorem3.2). Everything in[3]works except the boundedness of the Hessian of eigenfunctions around a critical point (since the eigenfunction is onlyC1,α amongM), which was used to prove (25). Here we avoid the use of the Hessian of eigenfunctions by using the com- parison theorem on the gradient. For the rest we follow step by step the work of Bakry–Qian[3]to get Theorem1.1.

This paper is organized as follows. In Section 2, the fundamentals in Finsler geometry is briefly introduced and the recent work of Ohta–Sturm is reviewed. We shall first prove the comparison theorem on the gradient and on the maxima of the eigenfunction and then Theorem1.1in Section 3.

2. Preliminaries on Finsler geometry

In this section we briefly recall the fundamentals of Finsler geometry, as well as the recent developments on the analysis of Finsler geometry by Ohta–Sturm[14–16]. For Finsler geometry, we refer to[1]and[19].

2.1. Finsler structure and Chern connection

LetMnbe a smooth, connectedn-dimensional manifold. A functionF:T M→ [0,∞)is called aFinsler structure if it satisfies the following properties:

(i) F isConT M\ {0};

(ii) F (x, t V )=t F (x, V )for all(x, y)T M and allt >0;

(iii) for every(x, V )T M\ {0}, the matrix gij(V ):= 2

∂Vi∂Vj

1 2F2

(x, V ) is positive definite.

Such a pair(Mn, F )is called aFinsler manifold. A Finsler structure is said to bereversibleif, in addition,F is even.

OtherwiseF is nonreversible. By aFinsler measure spacewe mean a triple(Mn, F, m)constituted with a smooth, connectedn-dimensional manifoldM, a Finsler structureF onMand a measuremonM.

Forx1, x2M, thedistance functionfromx1tox2is defined by

d(x1, x2):=inf

γ 1

0

F

˙ γ (t)

dt,

where the infimum is taken over all C1-curvesγ:[0,1] →M such that γ (0)=x1 and γ (1)=x2. Note that the distance function may not be symmetric unlessF is reversible. AC-curveγ:[0,1] →Mis called ageodesicif F (γ )˙ is constant and it is locally minimizing. ThediameterofMis defined by

(4)

d:= sup

x,yM

d(x, y).

Theforwardandbackward open ballsare defined by B+(x, r):=

yM:d(x, y) < r

, B(x, r):=

yM: d(y, x) < r . We denoteB±(x, r):=B+(x, r)B(x, r).

For every nonvanishing vector fieldV,gij(V )induces a Riemannian structuregV ofTxMvia gV(X, Y )=

n i,j=1

gij(V )XiYj, forX, YTxM.

In particular,gV(V , V )=F2(V ).

Letπ:T M\ {0} →Mthe projection map. The pull-back bundleπT Madmits a unique linear connection, which is theChern connection. The Chern connection is determined by the following structure equations, which characterize

“torsion freeness”:

DXVYDYVX= [X, Y] (7)

and “almostg-compatibility”

Z

gV(X, Y )

=gV

DZVX, Y +gV

X, DZVY +CV

DVZV , X, Y

(8) forVT M\ {0},X, Y, ZT M. Here

CV(X, Y, Z):=Cij k(V )XiYjZk=1 4

3F2

∂ViVjVk(V )XiYjZk

denotes the Cartan tensor andDXVY the covariant derivative with respect to reference vectorVT M\ {0}. We mention here that CV(V , X, Y )=0 due to the homogeneity ofF. In terms of the Chern connection, a geodesicγ satisfiesDγγ˙˙γ˙=0. For local computations in Finsler geometry, we refer to[19].

2.2. Hessian and Finsler-Laplacian

We shall introduce the Finsler-Laplacian on Finsler manifolds. First of all, we recall the notion of the Legendre transform.

Given a Finsler structureF onM, there is a natural dual normFon the cotangent spaceTM, which is defined by

F(x, ξ ):= sup

F (x,V )1

ξ(V ) for anyξTxM.

One can show thatFis also a Minkowski norm onTMand gij(ξ ):= 2

∂ξi∂ξj 1

2F2

(x, ξ )

is positive definite for every(x, ξ )TM\ {0}.

TheLegendre transformis defined by the mapl:TxMTxM:

l(V ):=gV(V ,·) forVTxM\ {0},

0 forV =0.

One can verify thatF (V )=F(l(V ))for anyVT M andgij(x, l(V ))is the inverse matrix ofgij(x, V ).

Let u:M→R be a smooth function onM andDu be its differential 1-form. Thegradient of u is defined as

u(x):=l1(Du(x))TxM. DenoteMu:= {Du=0}. Locally we can write in coordinates

u= n i,j=1

gij(x,u)∂u

∂xi

∂xj inMu.

(5)

TheHessianofuis defined by using Chern connection as

2u(X, Y )=gu

DXuu, Y

. (9)

One can show that ∇2u(X, Y ) is symmetric, see [22] and [16]. Indeed, using (7) and (8) and noticing that Cu(u, X, Y )=0, we have

gu

DXuu, Y

=X

gu(u, Y )

gu

u, DXuY

=XY (u)gu

u, DYuX+ [X, Y]

=Y X(u)+ [X, Y](u)gu

u, DYuX

− [X, Y](u)

=Y

gu(u, X)

gu

u, DYuX

=gu

DYuu, X .

In order to define a Laplacian on Finsler manifolds, we need a measurem(or a volume formdm) onM. From now on, we consider the Finsler measure space(M, F, m)equipped with a fixed smooth measurem. LetVT Mbe a smooth vector field onM. ThedivergenceofV with respect tomis defined by

divmV dm=d(Vdm),

whereVdmdenotes the inner product ofV with the volume formdm. In a local coordinate(xi), expressingdm= eΦdx1dx2· · ·dxn, we can write divmV as

divmV = n i=1

∂Vi

∂xi +Vi∂Φ

∂xi

.

A Laplacian, which is called theFinsler-Laplacian, can now be defined by mu:=divm(u).

We remark that the Finsler-Laplacian is better to be viewed in a weak sense that foruW1,2(M),

M

φmu dm= −

M

Dφ(u) dm forφCc(M).

The relationship betweenmuand∇2uis that mu+DΨ (u)=trgu

2u

= n i=1

2u(ei, ei),

whereΨ is defined bydm=eΨ (V )dVolgV and{ei}is an orthonormal basis ofTxM with respect togu. See e.g.

[22, Lemma 3.3].

Given a vector fieldV, theweighted Laplacianis defined on the weighted Riemannian manifold(M, gV, m)by Vmu:=divm

Vu , where

Vu:= ni,j=1gij(x, V )∂x∂u

i

∂xj forxMu,

0 forxM\Mu.

Similarly, the weighted Laplacian can be viewed in a weak sense that foruW1,2(M). We note thatmuu=mu.

2.3. Finsler manifolds with boundary

Assume that(M, F, m)is a Finsler measure space with boundary∂M, then we shall view∂M as a hypersurface embedded inM.∂Mis also a Finsler manifold with a Finsler structureF∂M induced byF. For anyx∂M, there exists exactly two unitnormal vectorsν, which are characterized by

Tx(∂M)=

VTxM: gν(ν, V )=0, gν(ν, ν)=1 .

(6)

Throughout this paper, we choose the normal vector that points outwardM. Note that, ifνis a normal vector,−νmay be not a normal vector unlessF is reversible.

The normal vectorνinduces a volume formdmνon∂Mfromdmby Vdm=gν(ν, V ) dmν, for allVT (∂M).

One can check that Stokes theorem holds (see[19, Theorem 2.4.2])

M

divm(V ) dm=

∂M

gν(ν, V ) dmν. We recall the convexity of the boundary ofM.

Thenormal curvatureΛν(V )atx∂Min a directionVTx(∂M)is defined by Λν(V )=gν

ν, Dγγ˙˙γ (0)˙

, (10)

whereγis the unique local geodesic for the Finsler structureF∂M on∂Minduced byF with the initial dataγ (0)=x andγ (0)˙ =V.

Mis said to hasconvex boundaryif for anyx∂M, the normal curvatureΛν atxis nonpositive in any directions VTx(∂M). We remark that the convexity ofMmeans thatDγγ˙˙γ (0)˙ lies at the same side ofTxMasM. Hence the choice of normal is not essential for the definition of convexity. (See Lemma 3.2 below.) There are several equivalent definitions of convexity, see for example[2]and[19].

2.4. Weighted Ricci curvature

The Ricci curvature of Finsler manifolds is defined as the trace of the flag curvature. Explicitly, given two linearly independent vectorsV , WTxM\ {0}, theflag curvatureis defined by

KV(V , W )= gV(RV(V , W )W, V ) gV(V , V )gV(W, W )gV(V , W )2, whereRV is theChern curvature(orRiemannian curvature):

RV(X, Y )Z=DVXDYVZDVYDXVZDV[X,Y]Z.

Then theRicci curvatureis defined by Ric(V ):=

n1

i=1

KV(V , ei),

wheree1, . . . , en1,F (V )V form an orthonormal basis ofTxMwith respect togV.

We recall the definition of the weighted Ricci curvature on Finsler manifolds, which was introduced by Ohta in[14], motivated by the work of Lott–Villani[12]and Sturm[20]on metric measure spaces.

Definition 2.1. (See[14].) Given a unit vector VTxM, letη:[−ε, ε] →M be the geodesic such thatη(0)˙ =V. Decomposemasm=eΨdvolη˙alongη, where volη˙ is the volume form ofgη˙as a Riemannian metric. Then

Ricn(V ):=

Ric(V )+η)(0) ifη)(0)=0,

−∞ otherwise;

RicN(V ):=Ric(V )+η)(0)η)(0)2

Nn , forN(n,), Ric(V ):=Ric(V )+η)(0).

Forc0 andN∈ [n,∞], define RicN(cV ):=c2RicN(V ).

Ohta proved in[14]that, forK∈R, the boundRicN(V )KF2(V )is equivalent to Lott–Villani and Sturm’sweak curvature-dimension condition CD(K, N ).

(7)

2.5. Bochner–Weitzenböck formula

The following Bochner–Weitzenböck type formula, established by Ohta–Sturm in[16], plays an important role in this paper.

Theorem 2.1.(See[16, Theorem 3.6].) GivenuWloc2,2(M)C1(M)withmuWloc1,2(M), we have

M

u

F2(x,u) 2

dm=

M

η

D(mu)(u)+Ric(u)+∇2u2

H S(u)

dm

as well as

M

u

F2(x,u) 2

dm

M

η

D(mu)(u)+RicN(u)+(mu)2 N

dm

for anyN ∈ [n,∞]and all nonnegative functionsηWc1,2(M)L(M). Here2u2H S(u)denotes the Hilbert–

Schmidt norm with respect togu.

Based on Bochner–Weitzenböck formula, a similar argument as Bakry–Qian[3, Theorem 6], leads to a refined inequality, which was referred to as an extended curvature-dimension inequality there. Another direct proof was also given in[21, Lemma 2.3].

Theorem 2.2.Assume that RicNK for someN ∈ [n,∞] and someK∈R. GivenuWloc2,2(M)C1(M)with muWloc1,2(M), we have

M

u

F2(x,u) 2

dm

M

η

D(mu)(u)+KF (u)2+(mu)2 N + N

N−1 mu

ND(F2(x,u))(u) 2F2(x,u)

2

dm (11)

for anyN∈ [n,∞]and all nonnegative functionsηWc1,2(M)L(M).

3. Proof of Theorem1.1

We first remark that a weak eigenfunctionuW1,2(M)of Finsler-Laplacian defined in (1) has regularity that uC1,α(M)W2,2(M)C(Mu)(see[6]).

Let us recall the 1-D modelsLK,Ndescribed in[3]. LetK∈RandN(1,∞].

(i) ForK >0 and 1< N <∞,LK,N is defined on(2K/(Nπ 1), π

2

K/(N1))by LK,N(v)(t)=v

K(N−1)tan K

N−1t

v;

(ii) ForK <0 and 1< N <∞,LK,N is defined on(−∞,0)∪(0,)by LK,N(v)(t)=v

K(N−1)coth

K N−1t

v and on(−∞,)by

LK,N(v)(t)=v

K(N−1)tanh

K N−1t

v;

(8)

(iii) ForK=0 and 1< N <∞,LK,Nis defined on(−∞,0)∪(0,)by LK,N(v)(t)=v+N−1

t v and on(−∞,)by

LK,N(v)(t)=v;

(iv) ForK=0 andN= ∞,LK,N is defined on(−∞,)by LK,N(v)(t)=vKt v;

(v) ForK=0 andN= ∞,LK,N is defined on(−∞,)by LK,N(v)(t)=vcv

for any constantc.

For convenience, we writeLK,N(v)(t)=vT (t )v. It is easy to check thatT=K+NT21. Denote byμK,N the invariant measure associated withLK,N, that is, a measure satisfyingb

aLK,N(v) dμK,N=0 forv(a)=v(b)=0.

For instance, in the case (i),K,N=cosN1( K

N1t ) dt.

The following gradient comparison theorem plays the most crucial role in the proof of our main theorem.

Theorem 3.1.Let(M, F, m)andλ1be as in Theorem1.1andube the eigenfunction. Letvbe a solution of the1-D model problem on some interval(a, b):

LK,N(v)= −λ1v, v(a)=v(b)=0, v>0. (12)

Assume that[minu,maxu] ⊂ [minv,maxv], then F

x,u(x) v

v1 u(x)

. (13)

Proof. First, since

Mu=0, minu <0 while maxu >0. We may assume that[minu,maxu] ⊂(minv,maxv)by multiplyingu by a constant 0< c <1. If we prove the result for this u, then letting c→1 implies the original statement.

Under the condition[minu,maxu] ⊂(minv,maxv),v1is smooth on a neighborhoodUof[minu,maxu].

ConsiderP (x)=ψ (u)(12F2(x,u)φ(u)), whereψ, φC(U )are two positive smooth functions to be deter- mined later. We first consider the case thatP attains its maximum atx0M, then study the case thatx0∂Mif∂M is not empty.

Case1.P attains its maximum atx0M.

Due to the lack of regularity ofu, we shall compute in the distributional sense. Letηbe any nonnegative function inWc1,2(M)L(M). We first compute

MDη(uP ) dm.

M

uP

dm= −

M

ψ

ψPψ φ

Dη(u)+ψ Dη

u 1

2F2(x,u)

dm

=

M

D ψ

ψPψ φ

η

(u)+ηD ψ

ψPψ φ

(u)

D(ψ η)

u 1

2F2(x,u)

+ηDψ

u 1

2F2(x,u)

dm :=I+II+III+IV.

(9)

By usingDu(u)=F2(x,u)=2(Pψ +φ)andmu= −λ1uin weak sense, we compute I=

M

λ1u ψ

ψPψ φ

η dm,

II=

M

η ψ

ψψ2 ψ2

Pψ φψφ

Du(u)+ψ

ψDP (u)

dm

=

M

η

2 ψ

ψψ2 ψ2

Pψ φψφ P

ψ +φ

+ψ

ψDP (u)

dm,

IV=

M

ηψ 1

ψDu

uP +

ψ ψ2P +φ

Du(u)

dm

=

M

2ηψ

ψ ψ2P+φ

P ψ +φ

+terms ofDP (u)

dm.

For the termIII, we apply the refined integral Bochner–Weitzenböck formula (11) to derive III

M

ψ η

D(mu)(u)+KF2+(mu)2

N + N

N−1 mu

ND(F2(x,u))(u) 2F2(x,u)

2 dm

=

M

ψ η

2(K−λ1) P

ψ+φ

+λ21u2

N + N

N−1 −λ1u

N

ψ ψ2P+φ

− 1

ψ F2DP (u) 2

dm

=

M

ψ η

2(K−λ1) P

ψ+φ

+ λ21u2 N−1+ N

N−1

ψ ψ2P +φ

2

+ 2 N−1λ1u

ψ ψ2P+φ

+terms ofDP (u)

dm.

Combining all we obtain

M

uP

dm

M

η 1

ψ

2ψ ψ

4− N

N−1 ψ2

ψ2

P2

+

ψ

ψ −2ψ2 ψ2

N+1 N−1

ψ

ψλu− 2N N−1

ψ

ψφ+2(K−λ1)−2φ

P +ψ

1

N−1λ21u2+N+1

N−1λ1+ N

N−1φ2+2(K−λ1−2φφ

+terms ofDP (u)

dm := −

M

a1P2+a2P +a3+terms ofDP (u)

dm. (14)

Therefore,

muP+terms ofDP (u)=a1P2+a2P +a3 (15)

holds in the distributional sense inM.

We claim that at the maximum pointx0ofP,

a1P2+a2P +a30. (16)

(10)

In fact, if not, then in a neighborhoodUofx0,a1P2+a2P+a3>0. It follows from (15) that the functionP is a strict subsolution to an elliptic operator inU. By maximum principle,P (x0) <maxUP, which contradicts the maximality ofP (x0).

It is interesting to see that the coefficientsai,i=1,2,3, coincide with that appeared in the Riemannian case (see e.g.[3, Lemma 1]). The next step is to choose suitable positive functionsψ andφsuch thata1, a2>0 everywhere anda3=0, which has already been done in[3]. For completeness, we sketch the main idea here.

Chooseφ(u)=12v(v1(u))2, wherevis a solution of 1-D problem (12). One can compute that φ(u)=v

v1(u)

, φ(u)=v v

v1(u) . Sett=v1(u)andu=v(t )then

a3(t)

ψ = 1

N−1λ21v2+N+1

N−1λ1vv+ N

N−1v2+(Kλ1)v2vv

= −v

vT v+λ1v + 1

N−1

vT v+λ1v

N v+T v+λ1v

=0.

Here we have used thatT satisfiesT=K+NT21. Fora1, a2, we introduce X(t )=λ1v(t )

v(t), ψ (u)=exp h v(t )

, f (t )= −h v(t )

v(t).

With these notations, we have f= −hv2+f (TX), v2

v1a1ψ=2f (T −X)N−2

N−1f2−2f:=2

Q1(f )f , a2=f

3N−1 N−1 T −2X

−2T N

N−1TX

f2f:=Q2(f )f.

We may now use Corollary 3 in [3], which says that there exists a bounded function f on [minu,maxu] ⊂ (minv,maxv)such thatf<min{Q1(f ), Q2(f )}.

In view of (16), we know that by our choice ofψandφ,P (x0)0, and henceP (x)0 for everyxM, which leads to (13).

Case2.∂M= ∅andx0∂M.

To handle this case, we need to define a new normal vector field on∂M, that is normal with respect to the Rieman- nian metricgu. To be more general, for everyXT M, there is a unique normal vector fieldνXsuch that

gXX, Y )=0 for anyYT (∂M), gXX, νX)=1, gν(ν, νX) >0. (17) A simple calculation shows that

gX(ν, νX) >0. (18)

Indeed, let νX=Z+ for some a∈R andZT (∂M). (17) tells that a >0. HencegX(ν, νX)=gX(1aXZ), νX)=1a>0.

The following lemma follows directly from the definition ofνandνX. Lemma 3.1.LetX, YT M. Then

gν(ν, Y )=0 ⇔ YT (∂M)gXX, Y )=0. 2 Define four sets

T±νM:=

YT M: gν(ν, Y ) >0(<0) and

(11)

T±νXM:=

YT M: gXX, Y ) >0(<0) .

We have the following simple but important observation, which may be familiar to expects.

Lemma 3.2.T+νM=T+νXM, TνM=TνXM.

Proof. We first claim that eitherT+νMT+νXM orT+νMTνXM. Otherwise, there are two vector fields Y1, Y2T+νM, such thatgXX, Y1) >0 andgXX, Y2) <0. Then by the continuity ofgXX,·)inT+νM, there existsYT+νMwithgXX, Y )=0, which meansgν(ν, Y )=0 from Lemma3.1. A contradiction. Taking into consideration thatνT+νXM, we see thatT+νMT+νXM. A similar argument implies thatT+νXMT+νM. The second equivalence follows in a similar way. 2

Return to the case whenP attains its maximum atx0∂M. Ifu(x0)=0, nothing needs to be proved. Thus we assumex0Mu. Recall thatPC(Mu). Sinceνupoints outward due to its definition, by taking normal derivative ofP with respect toνu, we have

DP (νu)(x0)0.

On one hand, the Neumann boundary condition∇uT (∂M)implies that guu,u)(x)=0,

or equivalently,

Du(νu)(x)=0 forx∂M.

Thus we have

DP (νu)(x0)=1 2ψ (u)

D

F2(u) u)

(x0). (19)

On the other hand, using (8) and the symmetry of∇2u, we have D

F2(u)

u)=D

gu(u,u)

u)=2gu

Dνu

u(u),u

=2gu

Duu(u), νu

. (20)

By the convexity of∂M, for anyXT (∂M),gν(ν, DXXX)0. In particular, setX= ∇u, we know that gν

ν, Duu(u)

0. (21)

It follows from Lemmas3.1and3.2that (21) is equivalent to gu

νu, Duu(u)

0. (22)

Combining (19), (20) and (22), we conclude thatDP (νu)(x0)≤0, and henceDP (νu)(x0)=0. The tangent derivatives ofP obviously vanish due to its maximality. Hence we have also

P (x0)=0.

Thus the proof for Case 1 works in this case. This finishes the proof of Theorem3.1. 2 Another ingredient is a comparison theorem for the maxima of the eigenfunctions.

Theorem 3.2.Let(M, F, m), λ1be as in Theorem1.1and1< N <∞. Letv=vK,Nbe a solution of the1-Dmodel problem on some interval(a, b) LK,Nv= −λ1v, with initial datav(a)= −1, v(a)=0, where

a=

2K/(Nπ 1) forK >0,

0 forK0

and b=b(a) be the first number after a withv(b)=0. DenotemK,N =vK,N(b)=max(v). Assume that λ1>

max{NKN1,0}andmin(u)= −1. ThenmaxumK,N.

(12)

Proof. We argue by contradiction. Suppose max(u) < mK,N. Then[minu,maxu] ⊂ [minv,maxv]. The condition λ1>max{NKN1,0}ensures that

b

π

2

K/(N1) forK >0,

∞ forK0,

which in turn ensures thatv>0 in(a, b). Hence we could apply Theorem3.1foruandv.

The same argument as Theorem 12 in[3]implies that the ratio R(c)=

{uc}u dm

{vc}v dμK,N

is increasing on[min(u),0]and decreasing on[0,max(u)]. Therefore, forc12, we have that m

{uc}

2

{uc}

|u|dm2R(0)

{vc}

|v|K,N2R(0)μK,N

{vc}

. (23)

Letc= −1+εforε >0 small. A simple calculation gives thatv(a)=λN1. Hence fort close toa,v(t)has positive lower and upper bound. Together withv(a)=0, we see thatv(t )v(a)C(ta)2. Thus ift∈ {v−1+ε}, then t(a, a+12). It follows that

μK,N

{v−1+ε}

μK,N

a, a+12

N/2. (24)

On the other hand, we shall prove that m

{u−1+ε}

m

B±

x0, Cε12

. (25)

Letx0Mbe such thatu(x0)= −1. For anyxB±(x0, δ)withδ small,u(x)is close to−1 ands:=v1(u(x))is close toa. Thus we see again from the upper bound ofvandv(a)=0 thatv(s)C(sa). Therefore, we have from Theorem3.1thatF (x,u(x))v(v1(u(x)))C(sa)andF (x,v1(u(x)))=(v1)(u(x))F (x,u(x))1.

In turn, we get sa=v1

u(x)

v1 u(x0)

F

˜ x,v1

u(x)˜ δδ, and

u(x)u(x0)+Fx,˜˜ ∇u(x)˜˜

δ−1+C(sa)δ−1+2,

for somex,˜ x˜˜∈B±(x0, δ). Letε=2, we concludeB±(x0, δ)⊂ {u−1+ε}, which implies (25).

Combining (23), (24) and (25), we see that there exists some constantC >0 such that m

B±(x0, r)

CrN. (26)

This will lead to a contradiction. In fact, since max(u) < mK,N andmK,N is continuous with respect to(K, N ), we also have that max(u) < mK,N for anyN> Nclose toN. Argued as before, we will obtain (26) withN instead ofN, i.e.

m

B±(x0, r)

CrN. (27)

However, the volume comparison theorem for Finsler manifolds under the assumption of lower bound forRicN (see [14, Theorem 7.3]), implies thatm(B±(x0, r))CrNforr >0 small. A contradiction to (27). The previous argument also works in the casex0∂M. The proof is completed. 2

Besides the comparison theorem on the gradient and maxima, in order to prove Theorem1.1, we also need some properties of the 1-D models, which has been extensively studied in[3]. We refer to[3]for the elementary properties, meanwhile we list two of them, one presents the full range of the maximum functionmK,N, the other reveals that the central interval has the lowest first Neumann eigenvalue.

Références

Documents relatifs

We want to follow the proof of the existence of Delaunay surfaces given in the previous section in order to prove the existence of a smooth family of normal graphs over the

Namely, he proved that if M is as in Lichnerowicz theorem, except that it has now a boundary such that its mean curvature with respect to the outward normal vector field is

Theorem 3.3 If the conformal group of a compact functionally smooth Finsler manifold is not precompact, then it is the canonical Riemannian sphere.. 3.3

Our main result consists in showing that among all sets Ω of R N symmetric about the origin, having prescribed Gaussian measure, μ 1 (Ω) is maximum if and only if Ω is the

Shen and Zhang [29] generalized Mo’s work to Finsler target manifolds, derived the first and second variation formulae, proved non-existence of non-constant stable harmonic maps

Riemannian manifolds of nonpositive curvature. We introduce asymptotic geodesics, the geodesic ray boundary and study visibility, introduced by.. P. Eberlein, and

Its main result is an asymptotically deterministic lower bound for the PE of the sum of a low compressibility approximation to the Stokes operator and a small scaled random

This result by Reilly is a generalization, to Riemannian mani- folds of class (R, 0) (R &gt; 0), of the well known theorem by Lichnerowicz [18 ] and Obata [21], which says that