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

On the Lower Bound of the Principal Eigenvalue of a Nonlinear Operator

Yucheng Tu Department of Mathematics, University of California, San Diego, La Jolla, CA 92093, USA [email protected]
Abstract.

We prove sharp lower bound estimates for the first nonzero eigenvalue of the non-linear elliptic diffusion operator LpL_{p} on a smooth metric measure space, without boundary or with a convex boundary and Neumann boundary condition, satisfying BE(κ,N)BE(\kappa,N) for κ0\kappa\neq 0. Our results extends the work of Koerber[12] for case κ=0\kappa=0 and Naber-Valtorta[10] for the pp-Laplacian.

Key words and phrases:
Eigenvalue estimates, generalized pp-Laplacian, Bakry-Émery curvature dimension, gradient comparison, and half-linear differential equations
2010 Mathematics Subject Classification:
35P15, 35P30

1. Introduction

Let MM be a compact manifold. The Laplacian operator on MM plays a key role in studying the geometry of MM, and one of the key quantity related to the Laplacian is its first nonzero eigenvalue λ1\lambda_{1}, also called the principal eigenvalue. There have been a lot of works on the estimate of λ1\lambda_{1} for the Neumann boundary value problem:

{Δu=λu on Muν=0 in M\begin{cases}\Delta u=-\lambda u\qquad\text{ on }M\\ \frac{\partial u}{\partial\nu}=0\qquad\text{ in }\partial M\end{cases}

In [11] Payne and Weinberger showed λ1π2/D2\lambda_{1}\geq\pi^{2}/D^{2} for the Laplacian on the convex subset of n\mathbb{R}^{n} with diameter DD. Later in [4] Cheeger gave a lower bound of λ1\lambda_{1} in terms of the isoperimetric constant on compact Riemannian manifolds. In [7], given that MM has nonnegative Ricci curvature, P. Li and Yau proved the lower bound π2/4D2\pi^{2}/4D^{2} by using gradient estimate. Later Zhong and Yang [14] used a barrier argument to prove the sharp lower bound π2/D2\pi^{2}/D^{2} for compact Riemannian manifold with nonnegative Ricci curvature. Afterwards Kroger in [6] used a gradient comparison technique to recover the result of Zhong and Yang, and furthermore he was able to deal with a negative Ricci lower bound case. It was in Bakry and Emery’s work [2] that the situation is generalized into a manifold with weighted volume measure, and the Laplacian is replaced by a general elliptic diffusion operator LL. By defining curvature-dimension condition, we can make sense of Ricci lower bound in the senario of smooth measure spaces. Later Bakry and Qian [3] used gradient comparison technique similar to Kroger to prove the sharp lower bound π2/D2\pi^{2}/D^{2} for λ1(L)\lambda_{1}(L) assuming MM to be BE(0,N)BE(0,N) for some N1N\geq 1. Later Andrews and Ni [1] recovered this result with a simple modulus of continuity method.

In recent years there is much attention to the nonlinear operator called pp-Laplacian Δp\Delta_{p}. In [12] he showed the sharp estimate λ1(p1)πppDp\lambda_{1}\geq(p-1)\frac{\pi_{p}^{p}}{D^{p}} for Riemannian manifolds with Ricci lower bound 0 and p>1p>1, where πp\pi_{p} is the half period of pp-sine function which will be defined later. The method used is a gradient comparison via Bochner formula for pp-Laplacian, and a fine ODE analysis of the one dimensional model solution. Later Naber and Valtorta [10] extended the result to the case Ricκ(n1)\text{Ric}\geq\kappa(n-1) for κ<0\kappa<0. The key improvement is the better understanding of the one dimensional model equation in the κ<0\kappa<0 case, which is considerably more complicated than non-negative case. Very recently Li-Wang in [8] and [9] used a modulus of continuity method to get the gradient comparison, in the case of drifted pp-Laplacian, which fits in the setting of a Bakry-Emery manifold with weight efe^{-f}, thus opening the possibility of studying the non-linear version of LL operator with drifted terms in metric measure spaces Satisfying BE(κ,N)BE(\kappa,N). For κ=0\kappa=0 case, [5] showed that (p1)πppDp(p-1)\frac{\pi_{p}^{p}}{D^{p}} is the sharp lower bound.

In this paper we follow the approaches of [3] and [10] to study the non-linear operator LpL_{p} on a compact manifold with possibly convex boundary(to be defined later). We extend the Theorem 1.1 of [5] to the κ0\kappa\neq 0 case, more precisely:

Theorem 1.1.

Let MM be compact and connected and LL be an elliptic diffusion operator with invariant measure mm. Assume that LL satisfies BE(κ,N)BE(\kappa,N) where κ0\kappa\neq 0. Let DD be diameter defined by the intrinsic distance metric on MM. Let uu be an eigenfunction associated with λ\lambda satisfying Neumann boundary condition if M\partial M\neq\emptyset, where λ\lambda is the first nonzero eigenvalue of LpL_{p}. Then denoting w(p1):=|w|p2ww^{(p-1)}:=|w|^{p-2}w, we have

  • (1)

    When κ>0\kappa>0, assuming further that Dπ/κD\leq\pi/\sqrt{\kappa}, we have a sharp comparison:

    λλD\lambda\geq\lambda_{D}

    where λD\lambda_{D} is the first nonzero eigenvalue of the following Neumann eigenvalue problem on [D/2,D/2][-D/2,D/2]:

    ddt[(w)(p1)](n1)κtan(κt)(w)(p1)+λw(p1)=0\frac{d}{dt}\big{[}(w^{\prime})^{(p-1)}\big{]}-(n-1)\sqrt{\kappa}\tan(\sqrt{\kappa}t)(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0
  • (2)

    When κ<0\kappa<0, we have a sharp comparison:

    λλD\lambda\geq\lambda_{D}

    where λD\lambda_{D} is the first nonzero eigenvalue of the following Neumann eigenvalue problem on [D/2,D/2][-D/2,D/2]:

    ddt[(w)(p1)]+(n1)κtanh(κt)(w)(p1)+λw(p1)=0\frac{d}{dt}\big{[}(w^{\prime})^{(p-1)}\big{]}+(n-1)\sqrt{-\kappa}\tanh(\sqrt{-\kappa t})(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0

The rest of this paper is devoted to the proof of Theorem 1.1. The basic structure is the following. In section 2 we introduce the setting and definitions related to the linear elliptic diffusion operator LL. In section 3 we define the non-linear operator LpL_{p} and its Neumann eigenvalue problem. In section 4 we use the Bochner formula to derive a useful estimate (Prop 4.2) which will be used in proving the gradient comparison theorem in section 5. In section 6 we study the associated three one-dimensional model equations and combined with section 7, we get maximum comparison between our model solutions and the solution to the Neumann eigenvalue problem. Finally combining the gradient, maximum and diameter comparison we prove the theorem in section 8.
Acknowledgment. The author would like to thank his advisor Professor Lei Ni for lots of encouragement and helpful suggestions, and Dr. Xiaolong Li for explaining his paper with Kui Wang [8] and [9] to him.

2. The geometry of elliptic diffusion operators

In this section we give some definitions which will be used later in our proof. First we introduce the elliptic diffusion operator, which is a natural generalization of a second order linear differential operator on a Riemannian manifold.

Definition 2.1.

A linear second order operator L:C(M)C(M)L:C^{\infty}(M)\to C^{\infty}(M) is called an elliptic diffusion operator if for any Φ:r\Phi:\mathbb{R}^{r}\to\mathbb{R} we have

L(Φ(f1,f2,,fr))=i=1riΦL(fi)+i,j=1rijΦΓ(fi,fj)L(\Phi(f_{1},f_{2},\dots,f_{r}))=\sum_{i=1}^{r}\partial_{i}\Phi L(f_{i})+\sum_{i,j=1}^{r}\partial_{i}\partial_{j}\Phi\Gamma(f_{i},f_{j})

and Γ(f,f)0\Gamma(f,f)\geq 0 with equality if and only if df=0df=0. Here Γ\Gamma is defined as

Γ(f,g):=12(L(fg)fLggLf).\Gamma(f,g):=\frac{1}{2}\big{(}L(fg)-fLg-gLf\big{)}.
Definition 2.2.

We say that a locally finite Borel measure mm is L-invariant if there is a generalized function ν\nu such that

MΓ(f,h)𝑑m=MfLh𝑑m+MfΓ(g,ν)𝑑m\int_{M}\Gamma(f,h)dm=-\int_{M}fLhdm+\int_{\partial M}f\Gamma(g,\nu)dm

holds for all smooth f,gf,g. ν\nu is called the outward normal function and is defined to be a set of pairs (νi,Ui)iI(\nu_{i},U_{i})_{i\in I} for a covering UiU_{i} of M\partial M such that νiC(Ui)\nu_{i}\in C^{\infty}(U_{i}) and Γ(νiνj,)|UiUj=0\Gamma(\nu_{i}-\nu_{j},\cdot)|_{U_{i}\cap U_{j}}=0.

Definition 2.3.

We define the intrinsic distance d:M×M[0,]d:M\times M\to[0,\infty] as:

d(x,y):=sup{f(x)f(y)|fC(M),Γ(f)1}d(x,y):=\sup\Big{\{}f(x)-f(y)|f\in C^{\infty}(M),\Gamma(f)\leq 1\Big{\}}

and the diameter of MM by D:=sup{d(x,y)|x,yM}D:=\sup\{d(x,y)|x,y\in M\}.

Definition 2.4.

For any ff, u,vC(M)u,v\in C^{\infty}(M), we define the Hessian by

Hf(u,v)=12(Γ(u,Γ(f,v))+Γ(v,Γ(f,u))Γ(f,Γ(u,v)))H_{f}(u,v)=\frac{1}{2}\Big{(}\Gamma(u,\Gamma(f,v))+\Gamma(v,\Gamma(f,u))-\Gamma(f,\Gamma(u,v))\Big{)}

and the Γ2\Gamma_{2}-operator by

Γ2(u,v)=12(L(Γ(u,v))Γ(u,Lv)Γ(v,Lu)).\Gamma_{2}(u,v)=\frac{1}{2}\Big{(}L(\Gamma(u,v))-\Gamma(u,Lv)-\Gamma(v,Lu)\Big{)}.
Definition 2.5.

We can define the NN-Ricci curvature as

RicN(f,f)(x)=inf{Γ2(ϕ,ϕ)(x)1N(Lϕ)2(x)|ϕC(M),Γ(ϕf)(x)=0}\text{Ric}_{N}(f,f)(x)=\inf\Big{\{}\Gamma_{2}(\phi,\phi)(x)-\frac{1}{N}(L\phi)^{2}(x)\Big{|}\phi\in C^{\infty}(M),\Gamma(\phi-f)(x)=0\Big{\}}

and let Ric=Ric\text{Ric}=\text{Ric}_{\infty}.

Let κ\kappa\in\mathbb{R} and N[1,]N\in[1,\infty], we say that LL satisfies BE(κ,N)BE(\kappa,N) condition if and only if

RicN(f,f)κΓ(f).\text{Ric}_{N}(f,f)\geq\kappa\Gamma(f).

for any fC(M)f\in C^{\infty}(M).

If MM has a boundary, the geometry of M\partial M also plays an important role in the eigenvalue estimate. We define the convexity of M\partial M as follows.

Definition 2.6.

Let ν\nu be the outward normal direction, and UMU\subset M be an open set, ϕ\phi,ηC(U)\eta\in C^{\infty}(U) such that Γ(ν,η)\Gamma(\nu,\eta) and Γ(ν,ϕ)=0\Gamma(\nu,\phi)=0 on UMU\cap\partial M. We define the second fundamental form on M\partial M by

II(ϕ,η)=Hϕ(η,ν)=12Γ(ν,Γ(η,ϕ)).II(\phi,\eta)=-H_{\phi}(\eta,\nu)=-\frac{1}{2}\Gamma(\nu,\Gamma(\eta,\phi)).

If for any ϕ\phi as above with Γ(ϕ)>0\Gamma(\phi)>0 on UMU\cap\partial M we have

II(ϕ,ϕ)0 on UMII(\phi,\phi)\leq 0\qquad\text{ on }U\cap\partial M

Then we say M\partial M is convex. If we have strict inequality then M\partial M is called strictly convex.

3. The generalized pp-Laplacian and its eigenvalue problem

Now we are going to work on the eigenvalue problem of the non-linear operator LpL_{p} derived from the previously defined LL. The generalized pp-Laplacian is defined by

Lpu(x)={Γ(u)p22(Lu+(p2)Hu(u,u)Γ(u)) if Γ(u)(x)0;0 otherwise. L_{p}u(x)=\begin{cases}\Gamma(u)^{\frac{p-2}{2}}\left(Lu+(p-2)\frac{H_{u}(u,u)}{\Gamma(u)}\right)&\text{ if }\Gamma(u)(x)\neq 0;\\ 0&\text{ otherwise. }\end{cases}

We also define

pu(η)={Γ(u)p22(Lη+(p2)Hη(u,u)Γ(u)) if Γ(u)(x)0;0 otherwise. {\mathcal{L}}_{p}^{u}(\eta)=\begin{cases}\Gamma(u)^{\frac{p-2}{2}}\left(L\eta+(p-2)\frac{H_{\eta}(u,u)}{\Gamma(u)}\right)&\text{ if }\Gamma(u)(x)\neq 0;\\ 0&\text{ otherwise. }\end{cases}

which is the linearization of LpL_{p}. Now we define the eigenvalue of LpL_{p}. If λ\lambda\in\mathbb{R} and uC2(M)u\in C^{2}(M) satisfies the Neumann boundary problem:

{Lpu=λu|u|p2 on MΓ(u,ν~)=0 on M\begin{cases}L_{p}u=-\lambda u|u|^{p-2}\qquad\text{ on }M^{\circ}\\ \Gamma(u,\tilde{\nu})=0\qquad\text{ on }\partial M\end{cases}

Then we call λ\lambda an eignevalue, and uu an eigenfunction of LpL_{p}, however, we may not always find a classical solution. To define the eigenfunction in a weak sense, we first use the invariance of mm to deduce the following integration-by-parts formula:

Lemma 3.1.

Let ϕC(M)\phi\in C^{\infty}(M) and uC2(M)u\in C^{2}(M) and Γ(u)>0\Gamma(u)>0 on supp(ϕ)\text{supp}(\phi). Then we have

MϕLpu𝑑m=MΓ(f)p22Γ(f,ϕ)𝑑m+MΓ(f,ν~)Γ(f)p22ϕ𝑑m\int_{M}\phi L_{p}udm=-\int_{M}\Gamma(f)^{\frac{p-2}{2}}\Gamma(f,\phi)dm+\int_{\partial M}\Gamma(f,\tilde{\nu})\Gamma(f)^{\frac{p-2}{2}}\phi dm

So we define the eigenvalue and eigenfunction by

Definition 3.1.

We say that λ\lambda is an eigenvalue of LpL_{p} if there is a uW1,p(M)u\in W^{1,p}(M) such that for any ϕC(M)\phi\in C^{\infty}(M) the following identity holds:

MΓ(u)p22Γ(u,ϕ)𝑑m=λMϕu|u|p2𝑑m\int_{M}\Gamma(u)^{\frac{p-2}{2}}\Gamma(u,\phi)dm=\lambda\int_{M}\phi u|u|^{p-2}dm

We have the following result concerning the regularity of principal eigenfunctions.

Lemma 3.2.

(Lemma 2.2 in [5]) If MM is a compact smooth Riemannian manifold with an elliptic diffusion operator LL and an LL-invariant measure mm. Then the principal eigenfunction is in C1,α(M)C^{1,\alpha}(M) for some α>0\alpha>0, and uu is smooth near points xMx\in M such that Γ(u)(x)0\Gamma(u)(x)\neq 0 and u(x)0u(x)\neq 0; for p<2p<2, uu is C3,αC^{3,\alpha}, and for p>2p>2, uu is C2,αC^{2,\alpha} near xx where Γ(u)(x)0\Gamma(u)(x)\neq 0 and u(x)=0u(x)=0.

4. Bochner formula

In this section, we will derive the Bochner formula and an estimate which is helpful to prove the gradient estimate in the next section.

Proposition 4.1 (Bochner formula).

Let uC1,α(M)u\in C^{1,\alpha}(M) be a first eigenfunction of LpL_{p}, and xMx\in M be a point such that Γ(u)(x)\Gamma(u)(x)\neq and u(x)0u(x)\neq 0. Then at xx we have the following formula:

1ppu(Γ(u)p2)=Γ(u)p22(Γ(Lpu,u)(p2)LpuAu)+Γ(u)p2(Γ2(u,u)+p(p2)Au2)\displaystyle\frac{1}{p}{\mathcal{L}}_{p}^{u}\left(\Gamma(u)^{\frac{p}{2}}\right)=\Gamma(u)^{\frac{p-2}{2}}\left(\Gamma(L_{p}u,u)-(p-2)L_{p}uA_{u}\right)+\Gamma(u)^{p-2}\left(\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}\right)
Proof.

c.f.[5], Lemma 3.1. \square

Proposition 4.2.

Suppose LL satisfies BE(κ,N)BE(\kappa,N) for some κ\kappa\in\mathbb{R} and N[1,]N\in[1,\infty]. Then for any nNn\geq N, we have for n(1,)n\in(1,\infty),

Γ(u)p2(Γ2(u,u)+p(p2)Au2)(Lpu)2n+nn1(Lpun(p1)Γ(u)p22Au)2+κΓ(u)p1\displaystyle\Gamma(u)^{p-2}\left(\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}\right)\geq\frac{(L_{p}u)^{2}}{n}+\frac{n}{n-1}\left(\frac{L_{p}u}{n}-(p-1)\Gamma(u)^{\frac{p-2}{2}}A_{u}\right)^{2}+\kappa\Gamma(u)^{p-1}

for n=n=\infty,

Γ(u)p2(Γ2(u,u)+p(p2)Au2)(p1)2Γ(u)p2Au2+κΓ(u)p1,\Gamma(u)^{\frac{p}{2}}\left(\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}\right)\geq(p-1)^{2}\Gamma(u)^{p-2}A_{u}^{2}+\kappa\Gamma(u)^{p-1},

for n=1n=1,

Γ(u)p2(Γ2(u,u)+p(p2)Au2)(Lpu)2+κΓ(u)p1\Gamma(u)^{\frac{p}{2}}\left(\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}\right)\geq(L_{p}u)^{2}+\kappa\Gamma(u)^{p-1}
Proof.

Following [5] Lemma 3.3, we can scale uu on both sides so that Γ(u)(x)=1\Gamma(u)(x)=1. We can assume n=Nn=N since B(κ,N)B(\kappa,N) implies B(κ,n)B(\kappa,n) for nNn\geq N. When n=1n=1, by the curvature-dimension inequality and Lu=trHu=AuLu=\text{tr}H_{u}=A_{u}, we get

Γ2(u,u)+p(p2)Au2κ+(Lu)2+p(p2)Au2=κ+(p1)2Au2=(Lpu)2+κ.\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}\geq\kappa+(Lu)^{2}+p(p-2)A_{u}^{2}=\kappa+(p-1)^{2}A_{u}^{2}=(L_{p}u)^{2}+\kappa.

When n=n=\infty, we have Γ2(u,u)κ+Au2\Gamma_{2}(u,u)\geq\kappa+A_{u}^{2}, therefore Γ2(u,u)+p(p2)Au2κ+(p1)2Au2\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}\geq\kappa+(p-1)^{2}A_{u}^{2}. Now if 1<n<1<n<\infty, for any vC(M)v\in C^{\infty}(M), by the curvature-dimension inequality we have

Γ2(v,v)κΓ(v)+1N(Lv)2\Gamma_{2}(v,v)\geq\kappa\Gamma(v)+\frac{1}{N}(Lv)^{2}

Now we consider a quadratic form B(v,v)=Γ2(v,v)κΓ(v)1N(Lv)2B(v,v)=\Gamma_{2}(v,v)-\kappa\Gamma(v)-\frac{1}{N}(Lv)^{2}, which is non-negative for any vC(M)v\in C^{\infty}(M). Let v=ϕ(u)v=\phi(u) where ϕC()\phi\in C^{\infty}(\mathbb{R}). Then by standard computations, together with the assumption Γ(u)=1\Gamma(u)=1, we have

Γ(ϕ(u))\displaystyle\Gamma(\phi(u)) =(ϕ)2,L(ϕ(u))=ϕLu+ϕ′′,\displaystyle=(\phi^{\prime})^{2},\qquad L(\phi(u))=\phi^{\prime}Lu+\phi^{\prime\prime},
Γ2(ϕ(u),ϕ(u))\displaystyle\Gamma_{2}(\phi(u),\phi(u)) =(ϕ)2Γ2(u,u)+2ϕϕ′′Au+(ϕ′′)2.\displaystyle=(\phi^{\prime})^{2}\Gamma_{2}(u,u)+2\phi^{\prime}\phi^{\prime\prime}A_{u}+(\phi^{\prime\prime})^{2}.

Then we get

B(ϕ(u),ϕ(u))\displaystyle B(\phi(u),\phi(u)) =Γ2(ϕ(u),ϕ(u))κΓ(ϕ(u))1N(L(ϕ(u)))2\displaystyle=\Gamma_{2}(\phi(u),\phi(u))-\kappa\Gamma(\phi(u))-\frac{1}{N}(L(\phi(u)))^{2}
=(ϕ)2Γ2(u,u)+2ϕϕ′′Au+(ϕ′′)2κ(ϕ)21N[ϕLu+ϕ′′]2\displaystyle=(\phi^{\prime})^{2}\Gamma_{2}(u,u)+2\phi^{\prime}\phi^{\prime\prime}A_{u}+(\phi^{\prime\prime})^{2}-\kappa(\phi^{\prime})^{2}-\frac{1}{N}\Big{[}\phi^{\prime}Lu+\phi^{\prime\prime}\Big{]}^{2}
=(ϕ)2B(u,u)+2ϕϕ′′(AuLuN)+N1N(ϕ′′)2\displaystyle=(\phi^{\prime})^{2}B(u,u)+2\phi^{\prime}\phi^{\prime\prime}(A_{u}-\frac{Lu}{N})+\frac{N-1}{N}(\phi^{\prime\prime})^{2}

Since B(ϕ(u),ϕ(u))0B(\phi(u),\phi(u))\geq 0 for any ϕ\phi, we have non-positive discriminant

B(u,u)N1N(AuLuN)20B(u,u)\frac{N-1}{N}-\Big{(}A_{u}-\frac{Lu}{N}\Big{)}^{2}\leq 0

Therefore we have

Γ2(u,u)+p(p2)Au2\displaystyle\Gamma_{2}(u,u)+p(p-2)A_{u}^{2}
=\displaystyle= κ+1N(Lu)2+B(u,u)+p(p2)Au2\displaystyle\kappa+\frac{1}{N}(Lu)^{2}+B(u,u)+p(p-2)A_{u}^{2}
\displaystyle\geq κ+1N(Lp(u)+(p2)Au)2+NN1(AuLp(u)+(p2)AuN)2+p(p2)Au2\displaystyle\kappa+\frac{1}{N}(L_{p}(u)+(p-2)A_{u})^{2}+\frac{N}{N-1}\Big{(}A_{u}-\frac{L_{p}(u)+(p-2)A_{u}}{N}\Big{)}^{2}+p(p-2)A_{u}^{2}
=\displaystyle= κ+1N(Lp(u))2+NN1(Lp(u)N(p1)Au)2\displaystyle\kappa+\frac{1}{N}(L_{p}(u))^{2}+\frac{N}{N-1}\Big{(}\frac{L_{p}(u)}{N}-(p-1)A_{u}\Big{)}^{2}

\square

5. Gradient Comparison Theorem and Its Applications

In this section we prove the gradient comparison theorem of the eigenfunction with the solution to the one-dimensional model.

Theorem 5.1.

Let uu be a weak solution of

Lpu=λu(p1)L_{p}u=-\lambda u^{(p-1)}

satisfying Neumann boundary condition if M\partial M\neq\emptyset, where λ\lambda is the first nonzero eigenvalue of LpL_{p}. Assume that LL satisfies BE(κ,N)BE(\kappa,N). Let Tκ:IT_{\kappa}:I\to\mathbb{R} be a function that satisfies Tκ=T2/(N1)+(N1)κT_{\kappa}^{\prime}=T^{2}/(N-1)+(N-1)\kappa, and w:[a,b]w:[a,b]\to\mathbb{R} be a solution of the following ODE:

{ddt[(w)(p1)]Tκ(w)(p1)+λw(p1)=0w(a)=1,w(a)=0\displaystyle\begin{cases}\frac{d}{dt}\big{[}(w^{\prime})^{(p-1)}\big{]}-T_{\kappa}(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0\\ w(a)=-1,\quad w^{\prime}(a)=0\end{cases} (5.1)

such that ww is strictly increasing on [a,b][a,b] and the range of uu is contained the range of ww. Then for all xMx\in M,

Γ(w1(u(x)))1.\Gamma(w^{-1}(u(x)))\leq 1.
Proof.

By scaling uu so that min(u)=1\min(u)=-1, we can assume that the range of uu is contained in the range of ww. By chain rule of Γ\Gamma what we need to show, equivalently, is

Γ(u)12(x)w(w1(u(x)))\Gamma(u)^{\frac{1}{2}}(x)\leq w^{\prime}(w^{-1}(u(x)))

for all xMx\in M. Since TκT_{\kappa} depends smoothly on κ\kappa, we will first prove that for any κ~<κ\tilde{\kappa}<\kappa, the gradient comparison holds when TκT_{\kappa} is replaced by Tκ~T_{\tilde{\kappa}} in (5.1), and then we can take κ~κ\tilde{\kappa}\to\kappa. This will give us a room to use proof by contradiction.

Now for c>0c>0 we denote ϕc=(cww1)p\phi_{c}=(cw^{\prime}\circ w^{-1})^{p}, and consider the function Zc:MZ_{c}:M\to\mathbb{R}

Zc(x)=Γ(u)p2(x)ϕc(u(x))Z_{c}(x)=\Gamma(u)^{\frac{p}{2}}(x)-\phi_{c}(u(x))

Assume for contradiction that Z1(x)>0Z_{1}(x)>0 for some xMx\in M. Let

c0=inf{c:Zc(x)>0 for some xM}c_{0}=\inf\{c:Z_{c}(x)>0\text{ for some }x\in M\}

By our definition of c0c_{0}, there is a x0Mx_{0}\in M such that Zc0(x0)=0Z_{c_{0}}(x_{0})=0 is the maximum of Zc0Z_{c_{0}}. Now we denote Zc0Z_{c_{0}} as ZZ, ϕc0\phi_{c_{0}} as ϕ\phi when there is no confusion. When x0x_{0} is in the interior of MM, this clearly implies the following equations:

Z(x0)=0\displaystyle Z(x_{0})=0 (5.2)
Γ(Z,u)(x0)=0\displaystyle\Gamma(Z,u)(x_{0})=0 (5.3)
1ppu(Z)(x0)0\displaystyle\frac{1}{p}{\mathcal{L}}_{p}^{u}(Z)(x_{0})\leq 0 (5.4)

If x0Mx_{0}\in\partial M, since Γ(u,ν~)=0\Gamma(u,\tilde{\nu})=0 by the Neumann boundary condition, we have that Γ(Z,u)=0\Gamma(Z,u)=0 at x0x_{0}. Since ZZ achieves maximum at x0x_{0} and M\partial M is convex, we have

0Γ(Z,ν~)\displaystyle 0\leq\Gamma(Z,\tilde{\nu}) =Γ(Γ(u)p2ϕ(u),ν~)=p2Γ(u)p22Γ(Γ(u),ν~)ϕ(u)Γ(u,ν~)\displaystyle=\Gamma(\Gamma(u)^{\frac{p}{2}}-\phi(u),\tilde{\nu})=\frac{p}{2}\Gamma(u)^{\frac{p-2}{2}}\Gamma(\Gamma(u),\tilde{\nu})-\phi^{\prime}(u)\Gamma(u,\tilde{\nu})
=pΓ(u)p22II(u,u)ϕ(u)00\displaystyle=-p\Gamma(u)^{\frac{p-2}{2}}II(u,u)-\phi^{\prime}(u)\cdot 0\leq 0

Therefore Γ(Z,ν~)(x0)=0\Gamma(Z,\tilde{\nu})(x_{0})=0. This implies that the second derivative of ZZ along the normal direction is nonpositive. On the other hand, the second derivatives along tangential directions are nonpositive, hence the ellipticity of pu{\mathcal{L}}_{p}^{u} implies that pu(Z)(x0)0{\mathcal{L}}_{p}^{u}(Z)(x_{0})\leq 0. Hence we comfirmed the three equations above for all xMx\in M.

From (5.2) we get

p2Γ(u)p22Γ(Γ(u),u)ϕ(u)Γ(u)=0\frac{p}{2}\Gamma(u)^{\frac{p-2}{2}}\Gamma(\Gamma(u),u)-\phi^{\prime}(u)\Gamma(u)=0

which implies ϕ(u)=pΓ(u)p22Au\phi^{\prime}(u)=p\Gamma(u)^{\frac{p-2}{2}}A_{u}. Now by calculation we have

1ppu(ϕ(u))=1p(ϕ(u)Lpu+(p1)ϕ′′(u)Γ(u)p2)\frac{1}{p}{\mathcal{L}}_{p}^{u}(\phi(u))=\frac{1}{p}\Big{(}\phi^{\prime}(u)L_{p}u+(p-1)\phi^{\prime\prime}(u)\Gamma(u)^{\frac{p}{2}}\Big{)}

By chain rule we have ϕ=p[(w)p2w′′]w1\phi^{\prime}=p\cdot\big{[}(w^{\prime})^{p-2}\cdot w^{\prime\prime}\big{]}\circ w^{-1}, and ϕ′′=p[(p2)(w′′)2+w′′′w](w)p4w1\phi^{\prime\prime}=p\Big{[}(p-2)(w^{\prime\prime})^{2}+w^{\prime\prime\prime}w^{\prime}\Big{]}\cdot(w^{\prime})^{p-4}\circ w^{-1}, and by differentiating the ODE satisfied by ww we have

(p1)(w)p3[(p2)(w′′)2+w′′′w]=Tκ~(w)p1+(p1)Tκ~w′′(w)p2λ(p1)wwp2(p-1)(w^{\prime})^{p-3}\Big{[}(p-2)(w^{\prime\prime})^{2}+w^{\prime\prime\prime}w^{\prime}\Big{]}=T_{\tilde{\kappa}}^{\prime}(w^{\prime})^{p-1}+(p-1)T_{\tilde{\kappa}}w^{\prime\prime}(w^{\prime})^{p-2}-\lambda(p-1)w^{\prime}w^{p-2}

Therefore

ϕ′′=pTκ~(w)p1+(p1)Tκ~w′′(w)p2λ(p1)wwp2ww1.\phi^{\prime\prime}=p\cdot\frac{T_{\tilde{\kappa}}^{\prime}(w^{\prime})^{p-1}+(p-1)T_{\tilde{\kappa}}w^{\prime\prime}(w^{\prime})^{p-2}-\lambda(p-1)w^{\prime}w^{p-2}}{w^{\prime}}\circ w^{-1}.

Now we evaluate the above expression at u(x0)u(x_{0}). Since ϕ(u)=p[(w)p2w′′]w1(u)=pΓ(u)p22Au\phi^{\prime}(u)=p\cdot\big{[}(w^{\prime})^{p-2}\cdot w^{\prime\prime}\big{]}\circ w^{-1}(u)=p\Gamma(u)^{\frac{p-2}{2}}A_{u}, and by (1) we have ϕ(u)=ww1(u)=Γ(u)p2\phi(u)=w^{\prime}\circ w^{-1}(u)=\Gamma(u)^{\frac{p}{2}}, we have

1ppu(ϕ(u))=λu(p1)Γ(u)p22Au+Tκ~Γ(u)p1+(p1)Tκ~Γ(u)2p32Auλ(p1)up2Γ(u)p2\displaystyle\frac{1}{p}{\mathcal{L}}_{p}^{u}(\phi(u))=-\lambda u^{(p-1)}\Gamma(u)^{\frac{p-2}{2}}A_{u}+T_{\tilde{\kappa}}^{\prime}\Gamma(u)^{p-1}+(p-1)T_{\tilde{\kappa}}\Gamma(u)^{\frac{2p-3}{2}}A_{u}-\lambda(p-1)u^{p-2}\Gamma(u)^{\frac{p}{2}} (5.5)

By the ODE evaluated at w1(u(x0))w^{-1}(u(x_{0})), we have

(p1)Γ(u)p22AuTκ~Γ(u)p12+λu(p1)=0(p-1)\Gamma(u)^{\frac{p-2}{2}}A_{u}-T_{\tilde{\kappa}}\Gamma(u)^{\frac{p-1}{2}}+\lambda u^{(p-1)}=0

Hence

(p1)Tκ~Γ(u)2p32Au=(p1)[(p1)Γ(u)p22Au+λu(p1)]Γ(u)p22Au.(p-1)T_{\tilde{\kappa}}\Gamma(u)^{\frac{2p-3}{2}}A_{u}=(p-1)\big{[}(p-1)\Gamma(u)^{\frac{p-2}{2}}A_{u}+\lambda u^{(p-1)}\big{]}\Gamma(u)^{\frac{p-2}{2}}A_{u}.

Plugging the above equation into the third term of (5.5)(5.5), we have

1ppu(ϕ(u))=λ(p2)u(p1)Γ(u)p22Au+Tκ~Γ(u)p1+(p1)2Γ(u)p2Au2λ(p1)up2Γ(u)p2\displaystyle\frac{1}{p}{\mathcal{L}}_{p}^{u}(\phi(u))=\lambda(p-2)u^{(p-1)}\Gamma(u)^{\frac{p-2}{2}}A_{u}+T_{\tilde{\kappa}}^{\prime}\Gamma(u)^{p-1}+(p-1)^{2}\Gamma(u)^{p-2}A_{u}^{2}-\lambda(p-1)u^{p-2}\Gamma(u)^{\frac{p}{2}}

We can assume that κ~\tilde{\kappa} has the same sign as κ\kappa since we can pick κ~\tilde{\kappa} sufficiently close to κ\kappa. Now we have Tκ~=Tκ~2/(n1)+κ~T_{\tilde{\kappa}}^{\prime}=T_{\tilde{\kappa}}^{2}/(n-1)+{\tilde{\kappa}} to rewrite the second term and finally get

1ppu(ϕ(u))=\displaystyle\frac{1}{p}{\mathcal{L}}_{p}^{u}(\phi(u))= λu(p1)Γ(u)p22Au+1n1[λu(p1)+(p1)Γ(u)p22Au]2+κ~Γ(u)p1\displaystyle-\lambda u^{(p-1)}\Gamma(u)^{\frac{p-2}{2}}A_{u}+\frac{1}{n-1}\Big{[}\lambda u^{(p-1)}+(p-1)\Gamma(u)^{\frac{p-2}{2}}A_{u}\Big{]}^{2}+{\tilde{\kappa}}\Gamma(u)^{p-1}
+(p1)2Γ(u)p2Au2+(p1)λu(p1)Γ(u)p22Auλ(p1)up2Γ(u)p2\displaystyle\quad+(p-1)^{2}\Gamma(u)^{p-2}A_{u}^{2}+(p-1)\lambda u^{(p-1)}\Gamma(u)^{\frac{p-2}{2}}A_{u}-\lambda(p-1)u^{p-2}\Gamma(u)^{\frac{p}{2}}
=\displaystyle= (p2)λu(p1)Γ(u)p22Auλ(p1)up2Γ(u)p2+κ~Γ(u)p1\displaystyle(p-2)\lambda u^{(p-1)}\Gamma(u)^{\frac{p-2}{2}}A_{u}-\lambda(p-1)u^{p-2}\Gamma(u)^{\frac{p}{2}}+{\tilde{\kappa}}\Gamma(u)^{p-1}
+nn1[λu(p1)n+(p1)Γ(u)p22Au]2+λ2u2p2n\displaystyle\quad+\frac{n}{n-1}\Big{[}\frac{\lambda u^{(p-1)}}{n}+(p-1)\Gamma(u)^{\frac{p-2}{2}}A_{u}\Big{]}^{2}+\frac{\lambda^{2}u^{2p-2}}{n}

By Proposition 4.1 and 4.2, together with the fact that Lpu=λu(p1)L_{p}u=-\lambda u^{(p-1)} we have

1ppu(Γ(u)p2)\displaystyle\frac{1}{p}{\mathcal{L}}_{p}^{u}\big{(}\Gamma(u)^{\frac{p}{2}}\big{)} =Γ(u)p22(Γ(Lpu,u)(p2)LpuAu)+Γ(u)p2(Γ2(u,u)+p(p2)Au2)\displaystyle=\Gamma(u)^{\frac{p-2}{2}}(\Gamma(L_{p}u,u)-(p-2)L_{p}uA_{u})+\Gamma(u)^{p-2}(\Gamma_{2}(u,u)+p(p-2)A_{u}^{2})
Γ(u)p22(λ(p1)u(p2)Γ(u)+λ(p2)u(p1)Au)+λ2u2p2n\displaystyle\geq\Gamma(u)^{\frac{p-2}{2}}\big{(}-\lambda(p-1)u^{(p-2)}\Gamma(u)+\lambda(p-2)u^{(p-1)}A_{u}\big{)}+\frac{\lambda^{2}u^{2p-2}}{n}
+nn1[λu(p1)n+(p1)Γ(u)p22Au]2+κΓ(u)p1\displaystyle\quad+\frac{n}{n-1}\Big{[}\frac{\lambda u^{(p-1)}}{n}+(p-1)\Gamma(u)^{\frac{p-2}{2}}A_{u}\Big{]}^{2}+\kappa\Gamma(u)^{p-1}

Hence in both cases, we have 1ppu(Γ(u)p2ϕ(u))(κκ~)Γ(u)p1>0\frac{1}{p}{\mathcal{L}}_{p}^{u}\big{(}\Gamma(u)^{\frac{p}{2}}-\phi(u)\big{)}\geq(\kappa-{\tilde{\kappa}})\Gamma(u)^{p-1}>0, which is a contradiction with the second derivative test. Therefore we conclude that Z10Z_{1}\leq 0 on MM, which implies our gradient comparison result. \square

Remark 5.2.

When 1<p<21<p<2 we know that uC2,αu\in C^{2,\alpha} near x0x_{0}, hence the Bochner formula can not be directly applied to x0x_{0}. In this case notice that uu does not vanish identically in a neighborhood of x0x_{0}, we can choose xx0x^{\prime}\to x_{0} with u(x)0u^{\prime}(x^{\prime})\neq 0. As we apply the Bochner formula at xx^{\prime}, The first term Γ(u)p22Γ(Lpu,u)=λΓ(u)p22Γ(u(p1),u)\Gamma(u)^{\frac{p-2}{2}}\Gamma(L_{p}u,u)=-\lambda\Gamma(u)^{\frac{p-2}{2}}\Gamma(u^{(p-1)},u) since uu is a eigenfunction. Now this diverging term will cancel with λ(p1)up2Γ(u)p2-\lambda(p-1)u^{p-2}\Gamma(u)^{\frac{p}{2}} in the expression of 1ppu(ϕ(u))\frac{1}{p}{\mathcal{L}}_{p}^{u}(\phi(u)), which makes it possible to define 1ppu(Γ(u)p2ϕ(u))(x0)\frac{1}{p}{\mathcal{L}}_{p}^{u}\big{(}\Gamma(u)^{\frac{p}{2}}-\phi(u)\big{)}(x_{0}) to be the limit of 1ppu(ϕ(u))(x)\frac{1}{p}{\mathcal{L}}_{p}^{u}(\phi(u))(x^{\prime}) as xx0x^{\prime}\to x_{0}. Therefore the previous proof still works when 1<p<21<p<2.

6. One-dimensional Models

In this section we will study the one-dimensional comparison model ODEs and discuss some fine properties of their solutions. Let n2n\geq 2, p>1p>1 be fixed.

6.1. The Model ODE

Let aa\in\mathbb{R}, p>1p>1 be fixed, we will consider the following form of initial value problem:

{ddt(w)(p1)T(t)(w)(p1)+λw(p1)=0w(a)=1,w(a)=0\displaystyle\begin{cases}\frac{d}{dt}(w^{\prime})^{(p-1)}-T(t)\cdot(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0\\ w(a)=-1,\quad w^{\prime}(a)=0\end{cases} (6.1)

where TT is defined over a subset of \mathbb{R}, to be specified according to the cases κ>0\kappa>0 or κ<0\kappa<0. To study this ODE we first define the pp-sin\sin and pp-cos\cos functions.

Definition 6.1.

For every p(1,)p\in(1,\infty), let πp\pi_{p} be defined by:

πp=11ds(1sp)1p=2πpsin(π/p)\pi_{p}=\int_{-1}^{1}\frac{ds}{(1-s^{p})^{\frac{1}{p}}}=\frac{2\pi}{p\sin(\pi/p)}

The C1C^{1} periodic function sinp:[1,1]\sin_{p}:\mathbb{R}\to[-1,1] is defined via the integral on [πp2,3πp2][-\frac{\pi_{p}}{2},\frac{3\pi_{p}}{2}] by

{t=0sinp(t)(1sp)1p𝑑s if t[πp2,πp2]sinp(t)=sinp(πpt) if t[πp2,3πp2]\begin{cases}t=\int_{0}^{\sin_{p}(t)}(1-s^{p})^{-\frac{1}{p}}ds\qquad&\text{ if }t\in\big{[}-\frac{\pi_{p}}{2},\frac{\pi_{p}}{2}\big{]}\\ \sin_{p}(t)=\sin_{p}(\pi_{p}-t)\qquad&\text{ if }t\in\big{[}\frac{\pi_{p}}{2},\frac{3\pi_{p}}{2}\big{]}\end{cases}

and we extend it to a periodic function on \mathbb{R}. Let cosp(t)=ddtsinp(t)\cos_{p}(t)=\frac{d}{dt}\sin_{p}(t), and we have the following identity which resembles the case of usual sin\sin and cos\cos:

|sinp(t)|p+|cosp(t)|p=1.|\sin_{p}(t)|^{p}+|\cos_{p}(t)|^{p}=1.

Let us use the Prüfer transformation to study the model equation (6.1).

Definition 6.2 (Prüfer transformation).

Let α=(λp1)1p\alpha=\big{(}\frac{\lambda}{p-1}\big{)}^{\frac{1}{p}}, then for some solution ww of the ODE, we define functions ee and ϕ\phi by

αw=esinp(ϕ)w=ecosp(ϕ).\alpha w=e\sin_{p}(\phi)\quad w^{\prime}=e\cos_{p}(\phi).

Standard calculation shows that ϕ\phi and ee satisfies the following first order systems:

{ϕ=αTp1cospp1(ϕ)sinp(ϕ)ϕ(a)=πp2\displaystyle\begin{cases}\phi^{\prime}=\alpha-\frac{T}{p-1}\cos_{p}^{p-1}(\phi)\sin_{p}(\phi)\\ \phi(a)=-\frac{\pi_{p}}{2}\end{cases} (6.2)
{ddtlog(e)=Tp1cospp(ϕ)e(a)=α\displaystyle\begin{cases}\frac{d}{dt}\log(e)=\frac{T}{p-1}\cos_{p}^{p}(\phi)\\ e(a)=\alpha\end{cases} (6.3)

6.2. Choice of TT in the case κ<0\kappa<0

When κ<0\kappa<0, we define 33 functions τi\tau_{i} on IiI_{i}\subset\mathbb{R}, i=1,2,3i=1,2,3:

  • (1)

    τ1(t)=sinh(κt)\tau_{1}(t)=\sinh(\sqrt{-\kappa}t)

  • (2)

    τ2(t)=exp(κt)\tau_{2}(t)=\exp(\sqrt{-\kappa}t)

  • (3)

    τ3(t)=cosh(κt)\tau_{3}(t)=\cosh(\sqrt{-\kappa}t)

and let μi=τin1\mu_{i}=\tau_{i}^{n-1}. Now we let Ti=μi/μiT_{i}=-\mu_{i}^{\prime}/\mu_{i} and we get:

  • (1)

    T1(t)=(n1)κ cotanh(κt)T_{1}(t)=-(n-1)\sqrt{-\kappa}\text{ cotanh}(\sqrt{-\kappa}t), defined on I1=(0,)I_{1}=(0,\infty);

  • (2)

    T2(t)=(n1)κT_{2}(t)=-(n-1)\sqrt{-\kappa}, defined on I2=I_{2}=\mathbb{R};

  • (3)

    T3(t)=(n1)κtanh(κt)T_{3}(t)=-(n-1)\sqrt{-\kappa}\tanh(\sqrt{-\kappa}t), defined on I3=I_{3}=\mathbb{R}.

6.3. Choice of TT in the case κ>0\kappa>0

When κ>0\kappa>0, let τ0(t)=cos(κt)\tau_{0}(t)=\cos(\sqrt{\kappa}t) and μ0=τ0n1\mu_{0}=\tau_{0}^{n-1}. Then T0=μ0/μ0T_{0}=\mu_{0}^{\prime}/\mu_{0} is defined as

T0(t)=(n1)κtan(κt)defined on I0=(π2κ,π2κ)T_{0}(t)=(n-1)\sqrt{\kappa}\tan(\sqrt{\kappa}t)\quad\text{defined on }I_{0}=(\frac{-\pi}{2\sqrt{\kappa}},\frac{\pi}{2\sqrt{\kappa}})

6.4. Fine analysis of the model equation (6.1)

The central question we need to address here is the existence of solution to (6.1) whose range matches the range of uu. Due to the normalization that minw=minu=1\min w=\min u=-1, we will consider the maximum of ww. For this purpose we introduce some notations. For aa\in\mathbb{R}, let wi,aw_{i,a} be the solution to the equation (6.1) with T=TiT=T_{i}, and b(i,a)b(i,a) be the first critical point of wi,aw_{i,a} after aa. If wi,a(t)>0w^{\prime}_{i,a}(t)>0 for t>at>a, then we say b(i,a)=b(i,a)=\infty. Also let δi,a=b(i,a)a\delta_{i,a}=b(i,a)-a and m(i,a)=wi,a(b(i,a))m(i,a)=w_{i,a}(b(i,a)). We shall prove the following statement in the current and next section:

Theorem 6.3.

Under the same setting as Theorem 1.1, let uu be an eigenfunction of LpL_{p} operator, normalized so that minu=1\min u=-1 and maxu1\max u\leq 1. Then we have the following existence results:

  • (i)

    (κ>0\kappa>0) There is some aI0a\in I_{0} and a solution w0,aw_{0,a} such that m(0,a)=maxum(0,a)=\max u.

  • (ii)

    (κ<0\kappa<0) There is some aa\in\mathbb{R}, i{1,2,3}i\in\{1,2,3\} and a solution wi,aw_{i,a} such that m(i,a)=maxum(i,a)=\max u.

To prove Theorem 6.3, first we establish the following existence and uniqueness of a solution to the model equation.

Proposition 6.1.

There is a unique solution to the initial value problem (6.1) with T=TiT=T_{i}, i=1,2,3i=1,2,3 in the following cases:(1) κ>0\kappa>0 and aI0{π/(2κ)}a\in I_{0}\cup\{-\pi/(2\sqrt{\kappa})\} and (2) κ<0\kappa<0 and aIi{0}a\in I_{i}\cup\{0\}.

Proof of Proposition 6.1.

In the case aIia\in I_{i} for i=0,1,2,3i=0,1,2,3 we obtain the existence and uniqueness result from the fact that TiT_{i} is a Lipschitz continuous function starting with aa. Hence we need to confirm the boundary cases only. When (1) κ>0\kappa>0 and a=π/(2κt)a=-\pi/(2\sqrt{\kappa}t) and (2) κ<0\kappa<0 and a=0a=0 for model T1T_{1}, we can use fixed-point theorem argument to prove the existence and uniqueness of the solution by slightly modifying the proof in [13], section 3. \square

Then we shall look at the case κ>0\kappa>0. In order to find aa such that w0,aw_{0,a} matches the maximum of uu, we use the continuous dependence of m(0,a)m(0,a) on aa. We need to show that

Proposition 6.2.

Fix α>0\alpha>0, n1n\geq 1 and κ>0\kappa>0. Then there always exists a unique a¯I0\bar{a}\in I_{0} such that the solution w3,a¯w_{3,-\bar{a}} is odd, and in particular, the maximum of ww restricted to [a¯,a¯][-\bar{a},\bar{a}] is 11.

The proof of Proposition 6.2 requires certain weaker estimate of λ\lambda. We define the first Neumann eigenvalue of the equation (6.1) on I0I_{0} to be

λ0:=inf{I0cosn1(t)|w|p𝑑tI0cosn1(t)|w|p𝑑t,wW1,p(I0){0}}\lambda_{0}:=\inf\bigg{\{}\frac{\int_{I_{0}}\cos^{n-1}(t)|w^{\prime}|^{p}dt}{\int_{I_{0}}\cos^{n-1}(t)|w|^{p}dt},\quad w\in W^{1,p}(I_{0})\setminus\{0\}\bigg{\}}

First we claim that

Lemma 6.1.

If λλ0\lambda\leq\lambda_{0}, κ>0\kappa>0, then equation (6.1) admits a odd solution ww such that w(t)>0w^{\prime}(t)>0 for all tI0t\in I_{0}.

Proof.

Consider the initial value problem starting with 0:

{ddt(w)(p1)T0(t)(w)(p1)+λw(p1)=0w(0)=0,w(0)>0\displaystyle\begin{cases}\frac{d}{dt}(w^{\prime})^{(p-1)}-T_{0}(t)\cdot(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0\\ w(0)=0,w^{\prime}(0)>0\end{cases} (6.4)

This problem admits a solution ww up to π/(2κ)\pi/(2\sqrt{\kappa}), the singularity of T0T_{0}, which can be extended to an odd solution on I0I_{0}. Now we claim that w(t)>0w^{\prime}(t)>0 for all tI0t\in I_{0}. Suppose for some t0(0,π/(2κ)t_{0}\in(0,\pi/(2\sqrt{\kappa}), w(t0)=0w^{\prime}(t_{0})=0. This is an eigenfunction corresponding to λ\lambda, hence by the monotonicity of first eigenvalue, we get λ>λ0\lambda>\lambda_{0}, which contradicts with our assumption that λλ0\lambda\leq\lambda_{0}. Therefore, we have w(t)>0w^{\prime}(t)>0 on I0I_{0}. \square

From Lemma 6.1, we can get a weaker bound on λ\lambda:

Lemma 6.2.

When κ>0\kappa>0, we have λ>λ0\lambda>\lambda_{0}.

Proof.

Suppose that on the contrary, λλ0\lambda\leq\lambda_{0}, then by Lemma 6.1, we get an odd function ww such that w>0w^{\prime}>0 on I0I_{0}. Suppose the first eigenfuntion uu is scaled so that 1=minumaxu1-1=\min u\leq\max u\leq 1. When ww is bounded, we can scale ww so that maxw=w(π/(2κ))=maxu\max w=w(\pi/(2\sqrt{\kappa}))=\max u. Pick x,yMx,y\in M such that u(x)=maxuu(x)=-\max u and u(y)=maxuu(y)=\max u. Then by the gradient comparison theorem,

DdistM(x,y)w1(u(y))w1(u(x))=πκD\geq\text{dist}_{M}(x,y)\geq w^{-1}(u(y))-w^{-1}(u(x))=\frac{\pi}{\sqrt{\kappa}}

which is a contradiction. When ww is unbounded, we can choose ck0c_{k}\searrow 0 and tkπ/(2κ)t_{k}\nearrow\pi/(2\sqrt{\kappa}) such that ckw(tk)=maxuc_{k}w(t_{k})=\max u. Again using the gradient comparison theorem, we can prove D2tkπ/κD\geq 2t_{k}\to\pi/\sqrt{\kappa}, which again gives a contradiction. Hence λ>λ0\lambda>\lambda_{0}. \square

Proof of Proposition 6.2.

Using the Prüfer Transformation, we consider the following initial value problem:

{ϕ(t)=αT0(t)p1cospp1(ϕ)sinp(ϕ)ϕ(0)=0\displaystyle\begin{cases}\phi^{\prime}(t)=\alpha-\frac{T_{0}(t)}{p-1}\cos_{p}^{p-1}(\phi)\sin_{p}(\phi)\\ \phi(0)=0\end{cases} (6.5)

Since λ>λ0\lambda>\lambda_{0}, equation 6.5 has a solution ϕ\phi such that ϕ(a^)=πp/2\phi(\hat{a})=\pi_{p}/2 for some a^<π/(2κ)\hat{a}<\pi/(2\sqrt{\kappa}). This implies that ee achieves maximum at a^\hat{a} from the equation satisfied by ee. Therefore we conclude that w=esinϕ/αw=e\sin\phi/\alpha achieves maximum at a^\hat{a}, and ww can be extended to an odd function on [a^,a^][-\hat{a},\hat{a}] such that w(a^)=0w^{\prime}(-\hat{a})=0, i.e. ww is a solution to the model equation (6.1). \square

By proposition 6.2, we know that m(0,a^)=1m(0,-\hat{a})=1. Now we show the continuous monotonicity of m(0,a)m(0,a), and first we need the following lemma to confirm continuity at the left endpoint:

Lemma 6.3.
limaπ/(2κ)m(0,a)=m(0,π/(2κ)).\lim_{a\to-\pi/(2\sqrt{\kappa})}m(0,a)=m(0,-\pi/(2\sqrt{\kappa})).
Proof of Lemma 6.3.

The idea of proof is from Proposition 1 of [3]. We will show that for any T<π/(2κ)T<\pi/(2\sqrt{\kappa}) and x(π/(2κ),T]x\in(-\pi/(2\sqrt{\kappa}),T], we have

limaπ/(2κ)wa(x)=wπ/(2κ)(x)andlimaπ/(2κ)wa(x)=wπ/(2κ)(x)\lim_{a\to-\pi/(2\sqrt{\kappa})}w_{a}(x)=w_{-\pi/(2\sqrt{\kappa})}(x)\quad\text{and}\quad\lim_{a\to-\pi/(2\sqrt{\kappa})}w_{a}^{\prime}(x)=w^{\prime}_{-\pi/(2\sqrt{\kappa})}(x)

We denote wπ/(2κ)(x)w_{-\pi/(2\sqrt{\kappa})}(x) by w(x)w(x). We consider the function Wa=wa(p)w(p)W_{a}=w_{a}^{(p)}-w^{(p)}, and know that the model equation 6.1 can be written as

(ρ[w(p)])+λρw(p)=0(\rho[w^{(p)}]^{\prime})^{\prime}+\lambda\rho w^{(p)}=0

where ρ(x)=cosn1(κx)\rho(x)=\cos^{n-1}(\sqrt{\kappa}x). Hence we have

(ρWa)+λρWa=0 on [a,T)(\rho W_{a}^{\prime})^{\prime}+\lambda\rho W_{a}=0\qquad\text{ on }[a,T)

Integrating the above equation over [a,x][a,x] we get

ρ(x)Wa(x)ρ(a)Wa(a)=λaxρ(t)Wa(t)𝑑t\rho(x)W_{a}^{\prime}(x)-\rho(a)W_{a}’(a)=-\lambda\int_{a}^{x}\rho(t)W_{a}(t)dt

Since Wa(a)=(w(p))(a)W_{a}^{\prime}(a)=-(w^{(p)})^{\prime}(a), we get

Wa(x)=(w(p))(a)ρ(a)ρ(x)λaxρ(t)ρ(x)Wa(t)𝑑t\displaystyle W_{a}^{\prime}(x)=-(w^{(p)})^{\prime}(a)\frac{\rho(a)}{\rho(x)}-\lambda\int_{a}^{x}\frac{\rho(t)}{\rho(x)}W_{a}(t)dt (6.6)

By another integration over [a,x][a,x], we have

Wa(x)=Wa(a)[w(p)](a)axρ(a)ρ(y)𝑑yλaxWa(t)txρ(t)ρ(y)𝑑y𝑑t\displaystyle W_{a}(x)=W_{a}(a)-[w^{(p)}]^{\prime}(a)\int_{a}^{x}\frac{\rho(a)}{\rho(y)}dy-\lambda\int_{a}^{x}W_{a}(t)\int_{t}^{x}\frac{\rho(t)}{\rho(y)}dydt (6.7)

We know that as TT is fixed,

|txρ(t)ρ(y)𝑑y|<C(n,T)\bigg{|}\int_{t}^{x}\frac{\rho(t)}{\rho(y)}dy\bigg{|}<C(n,T)

Since Wa(a)0W_{a}(a)\to 0 and [w(p)](a)0[w^{(p)}]^{\prime}(a)\to 0 as aπ/(2κ)a\to-\pi/(2\sqrt{\kappa}), the we have Wa(x)0W_{a}(x)\to 0 and Wa(x)0W_{a}^{\prime}(x)\to 0 as aπ/(2κ)a\to-\pi/(2\sqrt{\kappa}) by equation (6.6) and (6.7). \square

Proposition 6.3.

m(0,a)m(0,a) is an continuous monotonic function of aa on [π/(2κ),a^][-\pi/(2\sqrt{\kappa}),-\hat{a}].

Proof.

First we show that m(0,a)m(0,a) is an invertible function for a[π/(2κ),a^]a\in[-\pi/(2\sqrt{\kappa}),-\hat{a}]. Suppose there are aa and aa^{\prime} such that m(0,a)=m(0,a)m(0,a)=m(0,a^{\prime}). Then since w0,aw_{0,a} and w0,aw_{0,a^{\prime}} have same range and both are invertible functions on [a,b(a)][a,b(a)] and [a,b(a)][a^{\prime},b(a^{\prime})] respectively, by the gradient comparison theorem 5.1, we have

w0,aw0,a1=w0,aw0,a1w_{0,a}^{\prime}\circ w_{0,a}^{-1}=w_{0,a^{\prime}}^{\prime}\circ w_{0,a^{\prime}}^{-1}

and hence w0,a(x)=w0,a(xa+a)w_{0,a}(x)=w_{0,a^{\prime}}(x-a^{\prime}+a), i.e. identical under a translation. However, by the T0T_{0} model equation, this can only happen when a=aa=a^{\prime}. Therefore m(0,a)m(0,a) is invertible, and it is monotonic.
To see the continuity of m(0,a)m(0,a), note that when a>π/(2κ)a>-\pi/(2\sqrt{\kappa}), the continuous dependence of the solution on the initial value problem is automatic. When a=π/(2κ)a=-\pi/(2\sqrt{\kappa}), Lemma 6.3 shows that m(0,a)m(0,a) is continuous. \square

Now let us turn to the case κ<0\kappa<0, which is more delicate. We will get a similar result as Proposition 6.2:

Proposition 6.4 ([10] Proposition 6.1).

Fix α>0\alpha>0, n1n\geq 1 and κ<0\kappa<0. Then there always exists a unique a¯>0\bar{a}>0 such that the solution w3,a¯w_{3,-\bar{a}} is odd, and in particular, the maximum of ww restricted to [a¯,a¯][-\bar{a},\bar{a}] is 11.

By studying the equation of ϕ\phi one can show that there is a critical value α¯\bar{\alpha} at which the oscillatory behavior of ww changes. For the modeal T3T_{3}, we have

Lemma 6.4 ([10] Proposition 6.4).

There exists a limiting value α¯>0\bar{\alpha}>0 such that for α>α¯\alpha>\bar{\alpha} we have δ(3,a)<\delta(3,a)<\infty for every aa\in\mathbb{R}. For α<α¯\alpha<\bar{\alpha}, we have

limtϕ3,a(t)< for all a.\lim_{t\to\infty}\phi_{3,a}(t)<\infty\qquad\text{ for all }a\in\mathbb{R}.

for sufficiently large aa we have

πp2<limtϕ3,a(t)<0 and δ(3,a)=.-\frac{\pi_{p}}{2}<\lim_{t\to\infty}\phi_{3,a}(t)<0\qquad\text{ and }\delta(3,a)=\infty.

When α=α¯\alpha=\bar{\alpha}, we have limaδ(3,a)=.\lim_{a\to\infty}\delta(3,a)=\infty.

For model T1T_{1} we get the following result:

Lemma 6.5 ([10] Proposition 6.5).

There exist α¯>0\bar{\alpha}>0 such that when α>α¯\alpha>\bar{\alpha} then δ(1,a)<\delta(1,a)<\infty for all a[0,)a\in[0,\infty). If αα¯\alpha\leq\bar{\alpha} then ϕ1,a\phi_{1,a} has finite limit at infinity and δ(1,a)<\delta(1,a)<\infty for all a[0,)a\in[0,\infty).

Both cases α<α¯\alpha<\bar{\alpha} and αα¯\alpha\geq\bar{\alpha} need to be considered in proving the case (2) of Theorem 6.3. When α<α¯\alpha<\bar{\alpha} we can always use model T3T_{3} to produce the whole range comparison solutions ww, i.e. 0<maxw10<\max w\leq 1, and when αα¯\alpha\geq\bar{\alpha} we have restriction on the maximum value that uu can achieve. More precisely we have:

Lemma 6.6 ([10] Proposition 6.6).

Let αα¯\alpha\leq\bar{\alpha}. Then for each 0<maxu10<\max u\leq 1, there is an a[a¯,)a\in[-\bar{a},\infty) such that m(3,a)=maxum(3,a)=\max u.

We can also see that model T2T_{2} is translation invariant, hence for all a[0,)a\in[0,\infty), m(2,a)=m2m(2,a)=m_{2} is a constant. For model T1T_{1} and T3T_{3} we have

Lemma 6.7 ([10] Proposition 6.7).

If α>α¯\alpha>\bar{\alpha}, then m(3,a)m(3,a) is a decreasing function of aa, while m(1,a)m(1,a) is an increasing function of aa and

limam(3,a)=limam(1,a)=m2.\lim_{a\to\infty}m(3,a)=\lim_{a\to\infty}m(1,a)=m_{2}.

Combining the Proposition 6.4 and Lemmas above, we know that in the case κ<0\kappa<0, if m(1,0)maxu1m(1,0)\leq\max u\leq 1, there is always a model solution ww to T1,T2T_{1},T_{2} or T3T_{3} such that maxw=maxu\max w=\max u.

6.5. Diameter Comparison

In order to get the eigenvalue comparison with one-dimensional moder of the same diameter bound, we still need to understand how λD\lambda_{D} varies with the diameter. Again we will follow [10].

Definition 6.4.

We define the minimum diameter of the one-dimesional model associated with λ\lambda to be

δ¯i(λ)=min{δ(i,a)|i=0,1,2,3,aIi}\bar{\delta}_{i}(\lambda)=\min\{\delta(i,a)|i=0,1,2,3,a\in I_{i}\}

The following propositions deals with the lower bound of δ¯i(λ)\bar{\delta}_{i}(\lambda) for i=0,1,2,3i=0,1,2,3:

Proposition 6.5.

For i=0i=0 and any aI0a\in I_{0}, we have δ(0,a)>2a^\delta(0,a)>2\hat{a}, where a^<π/(2κ)\hat{a}<\pi/(2\sqrt{\kappa}) is such that w0,a^w_{0,-\hat{a}} is odd.

Proof.

The proof here is based on the symmetry and convexity of the model T0T_{0}. See [10] Proposition 8.4 for the proof. \square

For the case κ<0\kappa<0, we cite the following results from [10]:

Proposition 6.6 ([10], Proposition 8.2).

For i=1,2i=1,2 and any aIia\in I_{i}, we have δ(i,a)>πpα\delta(i,a)>\frac{\pi_{p}}{\alpha}.

Model 3 needs a little bit careful attention. For this one we notice first that there is always a¯>0\bar{a}>0 with an odd solution for initial data at a¯-\bar{a}. Namely w3,a¯w_{3,\bar{a}} is odd function with min 1-1 and max 11. This is a critical situation which minimizes the diameter DD given λ\lambda:

Proposition 6.7 ([10], Proposition 8.4).

For i=3i=3 and aa\in\mathbb{R}, we have

δ(3,a)δ(3,a¯)=2a¯\delta(3,a)\geq\delta(3,\bar{a})=2\bar{a}

and if aa¯a\neq-\bar{a}, the inequality is strict.

It is also easy to see from the ODE for ϕ\phi when i=3i=3 that, ϕ>α\phi^{\prime}>\alpha. Therefore δ(3,a¯)<πpα\delta(3,-\bar{a})<\frac{\pi_{p}}{\alpha}. Also from this we have δ(3,a¯)\delta(3,-\bar{a}) is strictly decreasing function of α\alpha, so as to λD\lambda_{D}. This means that δ¯(λ)\bar{\delta}(\lambda) is a strictly decreasing function. Thus if we see λ\lambda as a function of δ\delta, we also have the monotonicity: if δ1δ2\delta_{1}\leq\delta_{2}, we have

λ(δ1)λ(δ2).\lambda(\delta_{1})\geq\lambda(\delta_{2}).

7. Maximum of Eigenfunctions

In this section we are going to compare the maximum of the eigenfunctions uu and the model functions ww. First, we define a measure on the interval [a,b(a)][a,b(a)] which is essentially the pullback of the volume measure on MM by w1uw^{-1}\circ u. By the ODE satisfied by ww, w′′w^{\prime\prime} is positive before ww hits its first zero.

First we have a theorem which can be seen as a comparison between the model function and the eigenfunction.

Theorem 7.1.

(Theorem 34 [10]) Let uu and ww be as above and define

E(s):=exp(t0sw(p1)w(p1)𝑑t)asw(p1)𝑑μE(s):=-\exp{\bigg{(}\int_{t_{0}}^{s}\frac{w^{(p-1)}}{w^{\prime(p-1)}}dt\bigg{)}}\int_{a}^{s}w^{(p-1)}d\mu

then EE is increasing on (a,t0](a,t_{0}] and decreasing on [t0,b)[t_{0},b).

This result is equivalent to the following statement:

Theorem 7.2.

(Theorem 35,[10]) Under the hypothesis of Theorem 6.1 the function

E(s):=asw(p1)𝑑μasw(p1)tn1𝑑t=uw(s)u(p1)𝑑masw(p1)tn1𝑑tE(s):=\frac{\int_{a}^{s}w^{(p-1)}d\mu}{\int_{a}^{s}w^{(p-1)}t^{n-1}dt}=\frac{\int_{u\leq w(s)}u^{(p-1)}dm}{\int_{a}^{s}w^{(p-1)}t^{n-1}dt}

is increasing on (a,t0](a,t_{0}] and decreasing on [t0,b)[t_{0},b).

To prove the maximum comparison we study the volume of a small ball around the minimum of uu. By the gradient comparison we have the following:

Lemma 7.1.

For ϵ\epsilon sufficiently small, the set u1[1,1+ϵ)u^{-1}[-1,-1+\epsilon) contains a ball of radius w1(1+ϵ)aw^{-1}(-1+\epsilon)-a.

Now we can prove the maximum comparison, by combining Bishop-Gromov and the following:

Theorem 7.3.

Let nNn\geq N and n>1n>1. If uu is an eigenfunction satisfying minu=1=u(x0)\min u=-1=u(x_{0}) and maxum(1,0)=w1,0(b(1,0))\max u\leq m(1,0)=w_{1,0}(b(1,0)), then there exists a constant c>0c>0 such that for all rr sufficiently small, we have

m(Bx0(r))crn.m(B_{x_{0}}(r))\leq cr^{n}.
Proof.

To keep notations short, let w=w1,0w=w_{1,0}. Let ϵ\epsilon be small such that 1+ϵ<2p+1-1+\epsilon<-2^{-p+1}. Then we have u(p1)<12u^{(p-1)}<-\frac{1}{2} when u<1+ϵu<-1+\epsilon. Let t0t_{0} be the first zero of ww, then by Theorem 6.1 we have E(t)E(t0)E(t)\leq E(t_{0}). Therefore by Theorem 6.2 we get

m(Bx0(rϵ))Cu1+ϵu(p1)𝑑mCE(t0)aw1(1+ϵ)w(p1)tn1𝑑tCrϵnm(B_{x_{0}}(r_{\epsilon}))\leq C\int_{u\leq-1+\epsilon}u^{(p-1)}dm\leq CE(t_{0})\int_{a}^{w^{-1}(-1+\epsilon)}w^{(p-1)}t^{n-1}dt\leq C^{\prime}r_{\epsilon}^{n}

Since ϵ\epsilon can be arbitrarily small, we have the claim holds for rr sufficiently small. \square

Corollary 7.4.

Let nN>1n\geq N>1, and uu is an eigenfunction of LpL_{p} with minu=1\min u=-1 We have the following maximum comparison result:

  • (1)

    If κ>0\kappa>0, and w0,π/(2κ)w_{0,-\pi/(2\sqrt{\kappa})} is the corresponding eigenfunction. Then maxum(0,π/\max u\geq m(0,-\pi/ (2κ))(2\sqrt{\kappa})).

  • (2)

    If κ<0\kappa<0, and w1,a^w_{1,\hat{a}} is the corresponding eigenfunction. Then maxum(1,0)\max u\geq m(1,0).

Proof.

Let mm denotes m(0,π/(2κ))m(0,-\pi/(2\sqrt{\kappa})) or m(1,0)m(1,0) in either cases, and suppose that maxu<m\max u<m. Since mm is the least possible value among maxw\max w for all model solutions ww, by continuous dependence of the solution of model equation on nn, we can find n>nn^{\prime}>n so that maxu\max u is still less that the maximum of the correspoding model equation. Since BE(κ,n)BE(\kappa,n^{\prime}) is still satisfied, we have by Theorem 6.3, that m(Bx0(r))crnm(B_{x_{0}}(r))\leq cr^{n^{\prime}} for rr sufficiently small. However by Bishop-Gromov volume comparison we have m(Bx0(r))CrNm(B_{x_{0}}(r))\geq Cr^{N}. This is a contradiction since n>nNn^{\prime}>n\geq N. \square

8. Proof of Theorem 1.1

Now we can combine the gradient and maximum comparison, together with properties of the model equation to show the Theorem 1.1.

Theorem 8.1.

Let MM be compact and connected and LL be an elliptic diffusion operator with invariant measure mm. Assume that LL satisfies BE(κ,N)BE(\kappa,N) where κ0\kappa\neq 0. Let DD be diameter defined by the intrinsic distance metric on MM. Let uu be an eigenfunction associated with λ\lambda satisfying Neumann boundary condition if M\partial M\neq\emptyset, where λ\lambda is the first nonzero eigenvalue of LpL_{p}. Then denoting w(p1):=|w|p2ww^{(p-1)}:=|w|^{p-2}w, we have

  • (1)

    When κ>0\kappa>0, assuming further that Dπ/κD\leq\pi/\sqrt{\kappa}, we have a sharp comparison:

    λλD\lambda\geq\lambda_{D}

    where λD\lambda_{D} is the first nonzero eigenvalue of the following Neumann eigenvalue problem on [D/2,D/2][-D/2,D/2]:

    ddt[(w)(p1)](n1)κtan(κt)(w)(p1)+λw(p1)=0\frac{d}{dt}\big{[}(w^{\prime})^{(p-1)}\big{]}-(n-1)\sqrt{\kappa}\tan(\sqrt{\kappa}t)(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0
  • (2)

    When κ<0\kappa<0, we have a sharp comparison:

    λλD\lambda\geq\lambda_{D}

    where λD\lambda_{D} is the first nonzero eigenvalue of the following Neumann eigenvalue problem on [D/2,D/2][-D/2,D/2]:

    ddt[(w)(p1)]+(n1)κtanh(κt)(w)(p1)+λw(p1)=0\frac{d}{dt}\big{[}(w^{\prime})^{(p-1)}\big{]}+(n-1)\sqrt{-\kappa}\tanh(\sqrt{-\kappa t})(w^{\prime})^{(p-1)}+\lambda w^{(p-1)}=0
Proof.

We scale uu so that minu=1\min u=-1 and maxu1\max u\leq 1. By Proposition 6.3 we can find a model function wi,aw_{i,a} such that maxu=maxwi,a\max u=\max w_{i,a}. By the gradient comparison theorem, Γ(wi,a1u)1\Gamma(w_{i,a}^{-1}\circ u)\leq 1. Let xx and yy on MM be points where uu attains maximum and minimum, then we have

D|wi,a1u(x)wi,a1u(y)|=wi,a1(m(i,a))wi,a1(1)=δ(i,a,λ)δ(i,a¯)D\geq|w_{i,a}^{-1}\circ u(x)-w_{i,a}^{-1}\circ u(y)|=w_{i,a}^{-1}(m(i,a))-w_{i,a}^{-1}(-1)=\delta(i,a,\lambda)\geq\delta(i,\bar{a})

Therefore by the monotonicity of eigenvalue of the model equation, we have that

λλD.\lambda\geq\lambda_{D}.

To check the sharpness of this result when κ<0\kappa<0, we have the following examples: let

Mi=[D/2,D/2]×i1τ3Sn1M_{i}=[-D/2,D/2]\times_{i^{-1}\tau_{3}}S^{n-1}

be a warped product where Sn1S^{n-1} is the standard unit sphere, and τ3(t)=cosh(κt)\tau_{3}(t)=\cosh(\sqrt{-\kappa}t). If we consider LL being the classical Laplacian on MM, then standard computation shows that MiM_{i} has Ric(n1)κ\text{Ric}\geq-(n-1)\kappa and geodeiscially convex boundary. Hence it also satisfy the BE(κ,N)BE(\kappa,N) condition. If we take u(t,x)=w(t)u(t,x)=w(t) where ww is the solution to our one-dimensional model equation with λ=λD\lambda=\lambda_{D}. Since the diameter of MiM_{i} tends to dd as ii\to\infty, we see that the first eigenvalue on MiM_{i} converges to λd\lambda_{d}, which shows the sharpness of our lower bound when κ<0\kappa<0. For κ>0\kappa>0, the round sphere Sn1(πκ)S^{n-1}(\frac{\pi}{\sqrt{\kappa}}) serves as a model for sharp lower bound of λ\lambda. \square

References

  • [1] Ben Andrews and Lei Ni. Eigenvalue comparison on Bakry-Emery manifolds. Comm. Partial Differential Equations, 37(11):2081–2092, 2012.
  • [2] D Bakry and Michel Émery. Propaganda for γ\gamma2. From local times to global geometry, control and physics (Coventry, 1984/85), 150:39–46, 1986.
  • [3] Dominique Bakry and Zhongmin Qian. Some new results on eigenvectors via dimension, diameter, and Ricci curvature. Adv. Math., 155(1):98–153, 2000.
  • [4] Jeff Cheeger. A lower bound for the smallest eigenvalue of the laplacian. In Proceedings of the Princeton conference in honor of Professor S. Bochner, pages 195–199, 1969.
  • [5] Thomas Koerber. Sharp estimates for the principal eigenvalue of the p-operator. Calculus of Variations and Partial Differential Equations, 57(2):49, 2018.
  • [6] Pawel Kröger et al. On the spectral gap for compact manifolds. Journal of Differential Geometry, 36(2):315–330, 1992.
  • [7] Peter Li and Shing Tung Yau. Estimates of eigenvalues of a compact Riemannian manifold. In Geometry of the Laplace operator (Proc. Sympos. Pure Math., Univ. Hawaii, Honolulu, Hawaii, 1979), Proc. Sympos. Pure Math., XXXVI, pages 205–239. Amer. Math. Soc., Providence, R.I., 1980.
  • [8] Xiaolong Li and Kui Wang. Sharp lower bound for the first eigenvalue of the weighted pp-laplacian. arXiv:1910.02295, 2019.
  • [9] Xiaolong Li and Kui Wang. Sharp lower bound for the first eigenvalue of the weighted pp-laplacian ii. Math. Res. Lett, to appear, arXiv:1911.04596, 2019.
  • [10] Aaron Naber and Daniele Valtorta. Sharp estimates on the first eigenvalue of the pp-Laplacian with negative Ricci lower bound. Math. Z., 277(3-4):867–891, 2014.
  • [11] Lawrence E Payne and Hans F Weinberger. An optimal poincaré inequality for convex domains. Archive for Rational Mechanics and Analysis, 5(1):286–292, 1960.
  • [12] Daniele Valtorta. Sharp estimate on the first eigenvalue of the pp-Laplacian. Nonlinear Anal., 75(13):4974–4994, 2012.
  • [13] Wolfgang Walter. Sturm-liouville theory for the radial δp\delta_{p}-operator. Mathematische Zeitschrift, 227(1):175–185, 1998.
  • [14] Jia Qing Zhong and Hong Cang Yang. On the estimate of the first eigenvalue of a compact Riemannian manifold. Sci. Sinica Ser. A, 27(12):1265–1273, 1984.