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

Lower bounds for the first eigenvalue of pp-Laplacian on Kähler manifolds

Kui Wang School of Mathematical Sciences, Soochow University, Suzhou, 215006, China [email protected]  and  Shaoheng Zhang School of Mathematical Sciences, Soochow University, Suzhou, 215006, China [email protected]
Abstract.

We study the eigenvalue problem for the pp-Laplacian on Kähler manifolds. Our first result is a lower bound for the first nonzero eigenvalue of the pp-Laplacian on compact Kähler manifolds in terms of dimension, diameter, and lower bounds of holomorphic sectional curvature and orthogonal Ricci curvature for p(1,2]p\in(1,2]. Our second result is a sharp lower bound for the first Dirichlet eigenvalue of the pp-Laplacian on compact Kähler manifolds with smooth boundary for p(1,)p\in(1,\infty). Our results generalize corresponding results for the Laplace eigenvalues on Kähler manifolds proved in [14].

Key words and phrases:
Eigenvalue of pp-Laplacian, modulus of continuity, Kähler manifolds
2010 Mathematics Subject Classification:
35P15, 53C55
The research of the first author is supported by NSFC No.11601359

1. Introduction and Main Results

Let (Mm,g)(M^{m},g) be an mm-dimensional compact Riemannian manifold possibly with smooth boundary and the pp-Laplacian operator of the metric gg is defined by

Δpu:=div(|u|p2u)\Delta_{p}u:=\operatorname{div}(|\nabla u|^{p-2}\nabla u)

for p(1,)p\in(1,\infty). The pp-Laplacian eigenvalue equation is

(1.1) Δpu(x)=λ|u(x)|p2u(x),xM,-\Delta_{p}u(x)=\lambda|u(x)|^{p-2}u(x),\quad x\in M,

and λ\lambda\in\mathbb{R} is called a closed eigenvalue when M=\partial M=\emptyset, a Dirichlet eigenvalue when M\partial M\neq\emptyset and u(x)=0u(x)=0 on M\partial M, a Neumann eigenvalue when M\partial M\neq\emptyset and νu(x)=0\partial_{\nu}u(x)=0 on M\partial M. Where ν\nu denotes the outer unit normal to M\partial M.

The purpose of the present paper is to study the first nonzero eigenvalues of the pp-Laplacian on compact Kähler manifolds. In a recent paper [14], Li and the first author obtained a lower bound for the first nonzero closed and Neumann eigenvalue (when M\partial M\neq\emptyset) of the Laplacian on compact Kähler manifolds in terms of dimension, diameter, and lower bounds of holomorphic sectional curvature and orthogonal Ricci curvature. Precisely, they proved the following theorem.

Theorem 1.1 ([14]).

Let (Mm,g,J)\left(M^{m},g,J\right) be a compact Kähler manifold of complex dimension mm and diameter DD, whose holomorphic sectional curvature is bounded from below by 4κ14\kappa_{1} and orthogonal Ricci curvature is bounded from below by 2(m1)κ22(m-1)\kappa_{2} for some κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{R}. Let μ1,2(M)\mu_{1,2}(M) be the first nonzero eigenvalue of the Laplacian on MM (with Neumann boundary condition if MM has a strictly convex boundary). Then

μ1,2(M)μ¯1(m,κ1,κ2,D),\mu_{1,2}(M)\geq\bar{\mu}_{1}\left(m,\kappa_{1},\kappa_{2},D\right),

where μ¯1(m,κ1,κ2,D)\bar{\mu}_{1}\left(m,\kappa_{1},\kappa_{2},D\right) is the first Neumann eigenvalue of the one-dimensional eigenvalue problem

(1.2) φ′′(2(m1)Tκ2+T4κ1)φ=λφ\displaystyle\varphi^{\prime\prime}-\left(2(m-1)T_{\kappa_{2}}+T_{4\kappa_{1}}\right)\varphi^{\prime}=-\lambda\varphi

on [D/2,D/2][-D/2,D/2]. Here and in the rest of this paper, we denote the function TκT_{\kappa} for κ\kappa\in\mathbb{R} by

(1.3) Tκ(t)={κtan(κt),κ>0,0,κ=0,κtanh(κt),κ<0.T_{\kappa}(t)=\begin{cases}\sqrt{\kappa}\tan(\sqrt{\kappa}t),&\kappa>0,\\ 0,&\kappa=0,\\ -\sqrt{-\kappa}\tanh(\sqrt{-\kappa}t),&\kappa<0.\end{cases}

On Riemannian manifolds, the above theorem known as Zhong-Yang estimate is proved by Zhong and Yang [27] for the case Ric0\operatorname{Ric}\geq 0, and by Kröger [9] for the case Ric(n1)κ\operatorname{Ric}\geq(n-1)\kappa for general κ\kappa\in\mathbb{R}. These works use the gradient estimates method initiated by Li [10] and Li-Yau [12]. In 2013, Andrews and Clutterbuck [4] gave a simple proof via the modulus of continuity estimates for the solutions to the heat equation, see also [6] for a coupling method, and [26] for an elliptic proof based on [1] and [20]. Recently, Valtorta [25] and Naber-Valtorta [19] established the Zhong-Yang estimate for the first nonzero eigenvalue for pp-Laplacian on Riemannian manifolds.

On compact Kähler manifolds, Li and the first author proved a lower bound of the first nonzero eigenvalue of Laplacian in terms of geometric data (see Theorem 1.1), and Rutkowski and Seto [22] established an explicit lower bound by studying the one-dimensional ODE (1.2). On noncompact Kähler manifolds, Li-Wang [11] and Munteanu [18] obtained upper bounds of the first Laplace eigenvalue under conditions of bisectional curvature and Ricci curvature respectively. On closed Kähler manifolds with positive Ricci curvature bound from below, Lichnerowicz [17] obtained a optimal lower bound of the first nonzero eigenvalue of Laplacian by using Reilly formula, and generalized by Blacker and Seto [5] to the first nonzero eigenvalue of pp-Laplacian for p2p\geq 2. It is thus a natural question to study the first nonzero eigenvalue of pp-Laplacian on Kähler manifolds for p(1,2]p\in(1,2]. Regarding this, we prove

Theorem 1.2.

Let (Mm,g,J)(M^{m},g,J) be a compact Kähler manifold of complex dimension mm and diameter DD, whose holomorphic sectional curvature is bounded from below by 4κ14\kappa_{1} and orthogonal Ricci curvature is bounded from below by 2(m1)κ22(m-1)\kappa_{2} for some κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{R}. Let μ1,p(M)\mu_{1,p}(M) be the first nonzero eigenvalue of the pp-Laplacian on M (with Neumann boundary condition if MM has a strictly convex boundary). Assume 1<p21<p\leq 2, then

(1.4) μ1,p(M)μ¯1(m,p,κ1,κ2,D),\displaystyle\mu_{1,p}(M)\geq\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D),

where μ¯1(m,p,κ1,κ2,D)\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D) is the first nonzero Neumann eigenvalue of the one-dimensional eigenvalue problem

(1.5) (p1)|φ|p2φ′′(2(m1)Tκ2+T4κ1)|φ|p2φ=μ|φ|p2φ(p-1)|\varphi^{\prime}|^{p-2}\varphi^{\prime\prime}-(2(m-1)T_{\kappa_{2}}+T_{4\kappa_{1}})|\varphi^{\prime}|^{p-2}\varphi^{\prime}=-\mu|\varphi|^{p-2}\varphi

on interval [D/2,D/2][-D/2,D/2].

Note that if κ1=κ2=0\kappa_{1}=\kappa_{2}=0, ODE (1.5) can be solved by pp-trigonometric functions, then as a direct consequence of Theorem 1.2, we have

Corollary 1.3.

With the same assumptions as in Theorem 1.2, and assume κ1=κ2=0\kappa_{1}=\kappa_{2}=0. Then

(1.6) μ1,p(M)(p1)(πpD)p,\mu_{1,p}(M)\geq(p-1)(\frac{\pi_{p}}{D})^{p},

where πp=2πpsin(π/p)\displaystyle\pi_{p}=\frac{2\pi}{p\sin(\pi/p)}.

Now let us turn to lower bounds for Dirichlet eigenvalues on compact Kähler manifolds with smooth boundary. For κ,Λ\kappa,\Lambda\in\mathbb{R}, denote by Cκ,ΛC_{\kappa,\Lambda} the unique solution of the initial value problem

(1.7) {ϕ′′(t)+κϕ(t)=0,ϕ(0)=1,ϕ(0)=Λ,\left\{\begin{aligned} &\phi^{\prime\prime}(t)+\kappa\phi(t)=0,\\ &\phi(0)=1,\ \phi^{\prime}(0)=-\Lambda,\end{aligned}\right.

and denote Tκ,ΛT_{\kappa,\Lambda} for κ,Λ\kappa,\Lambda\in\mathbb{R} by

(1.8) Tκ,Λ(t):=Cκ,Λ(t)Cκ,Λ(t).T_{\kappa,\Lambda}(t):=-\frac{C_{\kappa,\Lambda}^{\prime}(t)}{C_{\kappa,\Lambda}(t)}.

For the first Dirichlet eigenvalue on compact Kähler manifolds, Li and the first author [14, Theorem 1.4] obtained a lower bound of the first Dirichlet eigenvalue of Laplacian using Barta’s inequality, and Blacker-Seto [5] proved a lower bound of the first Dirichlet eigenvalue of pp-Laplacian for p2p\geq 2 via the pp-Reilly formula. Our second main result is the following pp-Laplacian analogue of the lower bounds for the first Dirichlet eigenvalue on Kähler manifolds for 1<p<1<p<\infty.

Theorem 1.4.

Let (Mm,g,J)(M^{m},g,J) be a compact Kähler manifold with smooth boundary M\partial M. Suppose that the holomorphic sectional curvature is bounded from below by 4κ14\kappa_{1} and the orthogonal Ricci curvature is bounded from below by 2(m1)κ22(m-1)\kappa_{2} for κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{R}, and the second fundamental form on M\partial M is bounded from below by Λ\Lambda\in\mathbb{R}. Let λ1,p(M)\lambda_{1,p}(M) be the first Dirichlet eigenvalue of the pp-Laplacian on MM. Assume 1<p<1<p<\infty, then

(1.9) λ1,p(M)λ¯1(m,p,κ1,κ2,D),\displaystyle\lambda_{1,p}(M)\geq\bar{\lambda}_{1}(m,p,\kappa_{1},\kappa_{2},D),

where λ¯1(m,p,κ1,κ2,D)\bar{\lambda}_{1}(m,p,\kappa_{1},\kappa_{2},D) is the first eigenvalue of the one-dimensional eigenvalue problem

(1.10) (p1)|φ|p2φ′′(2(m1)Tκ2,Λ+T4κ1,Λ)|φ|p2φ=λ|φ|p2φ\displaystyle(p-1)|\varphi^{\prime}|^{p-2}\varphi^{\prime\prime}-\big{(}2(m-1)T_{\kappa_{2},\Lambda}+T_{4\kappa_{1},\Lambda}\big{)}|\varphi^{\prime}|^{p-2}\varphi^{\prime}=-\lambda|\varphi|^{p-2}\varphi

on [0,D/2][0,D/2] with boundary conditions φ(0)=0\varphi(0)=0 and φ(R)=0\varphi^{\prime}(R)=0.

The proof of Theorem 1.4 relies on a comparison theorem for the second derivatives of distance to the boundary proved in [14, Section 6] and Barta’s inequality. In the Riemannian case, lower bounds for the first Dirichlet eigenvalue were proved by Li-Yau [12] and by Kause [8]. It remains an interesting question that whether or not the same result as Theorem 1.2 holds for p>2p>2 and the result is sharp. We plan to return to this in the future. The rest of the paper is organized as follows. In Section 2, we recall the definitions of curvatures of Kähler manifolds and modulus of continuity of real functions. Sections 3 and 4 are devoted to proving Theorem 1.2 and Theorem 1.4.

2. Preliminary

2.1. Curvatures of Kähler Manifolds

Let (Mm,g,J)(M^{m},g,J) be a Kähler manifold with complex dimension mm (real dimension is n=2mn=2m). A plane σTpM\sigma\subset T_{p}M is said to be holomorphic if it is invariant by the complex structure tensor JJ. The restriction of the sectional curvature to holomorphic planes is called the holomorphic sectional curvature, denoted by HH. In other words, if σ\sigma is a holomorphic plane spanned by XX and JXJX, then the holomorphic sectional curvature of σ\sigma is defined by

H(σ):=H(X)=R(X,JX,X,JX)|X|4.H(\sigma):=H(X)=\frac{R(X,JX,X,JX)}{|X|^{4}}.

We say the holomorphic sectional curvature is bounded from below by κ\kappa, if H(σ)κH(\sigma)\geq\kappa for all holomorphic planes σTpM\sigma\subset T_{p}M and all pMp\in M. The orthogonal Ricci curvature, denoted by Ric\operatorname{Ric}^{\perp}, is defined for any XTpMX\in T_{p}M by

Ric(X,X):=Ric(X,X)H(X)|X|2.\operatorname{Ric}^{\perp}(X,X):=\operatorname{Ric}(X,X)-H(X)|X|^{2}.

We say the orthogonal Ricci curvature is bounded from below by κ\kappa\in\mathbb{R}, if Ric(X,X)κg(X,X),XTpM,pM\operatorname{Ric}^{\perp}(X,X)\geq\kappa g(X,X),\forall X\in T_{p}M,\forall p\in M,

Remark 2.1.

If MM is a complete Kähler manifold with holomorphic sectional curvature satisfies Hκ1H\geq\kappa_{1} for some positive κ1\kappa_{1}, Tsukamoto [24] proved that the diameter of MM is bounded from above by π/κ1\pi/\sqrt{\kappa_{1}}. If MM is a complete Kähler manifold with orthogonal Ricci curvature satisfies Ric2(m1)κ2\operatorname{Ric}^{\perp}\geq 2(m-1)\kappa_{2} for some positive κ2\kappa_{2}, Ni and Zheng [21] proved that the diameter of MM is bounded from above by π/κ2\pi/\sqrt{\kappa_{2}}.

2.2. Modulus of Continuity

Let uu be a continuous function on a metric space (X,d)(X,d), define the modulus of continuity ω\omega of uu by

ω(s):=sup{u(x)u(y)2|d(x,y)=2s,x,yX}.\omega(s):=\sup\big{\{}\frac{u(x)-u(y)}{2}|\ d(x,y)=2s,x,y\in X\big{\}}.

Recently, Andrews and Clutterbuck [1, 2, 3, 4] investigated how the modulus of continuity of solutions to parabolic differential equations evolves, and they proved for a large class of parabolic equation, the modulus of continuity of the solution is a viscosity solution of the associated one-dimensional equations. By using the modulus of continuity, Andrews and Clutterbuck obtained the sharp lower bound on the fundamental gap for Schrödinger operators [3]. This technique is called modulus of continuity estimate, and is widely used to study the lower bound of the first nonzero eigenvalue in terms of geometric data such as the diameter and dimension of the manifold, see [7, 16, 23] and so on.

3. The First Nonzero Eigenvalue

In this section, we shall use the method of modulus of continuity estimates to prove Theorem 1.2. The proof presented below is a modification of the argument outlined in the survey by Andrews [1, Section 8] for the case of Riemannian manifolds. Let φ(s,t)C2,1([0,a]×[0,))\varphi(s,t)\in C^{2,1}([0,a]\times[0,\infty)), and denote by

(3.1) 𝔏φ:=(p1)|φ|p2φ′′(2(m1)Tκ2+T4κ1)|φ|p2φ,\displaystyle\mathfrak{L}\varphi:=(p-1)|\varphi^{\prime}|^{p-2}\varphi^{\prime\prime}-(2(m-1)T_{\kappa_{2}}+T_{4\kappa_{1}})|\varphi^{\prime}|^{p-2}\varphi^{\prime},

where we denote kφsk\frac{\partial^{k}\varphi}{\partial s^{k}} by φ(k)\varphi^{(k)} for short. We first prove the following modulus of continuity estimates for solutions to a nonlinear parabolic equation on Kähler manifolds.

Theorem 3.1.

Let (Mm,g,J)(M^{m},g,J) be a compact Kähler manifold with diameter DD whose holomorphic sectional curvature is bounded from below by 4κ14\kappa_{1} and the orthogonal Ricci curvature is bounded from below by 2(m1)κ22(m-1)\kappa_{2} for some κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{R}. Let v:M×[0,T)v:M\times[0,T)\rightarrow\mathbb{R} be a solution of

(3.2) vt=|Δpv|p2p1Δpvv_{t}=|\Delta_{p}v|^{-\frac{p-2}{p-1}}\Delta_{p}v

(with Neumann boundary condition if MM has a strictly convex boundary). Suppose 1<p21<p\leq 2 and φ(s,t):[0,D2]×[0,)\varphi(s,t):[0,\frac{D}{2}]\times[0,\infty)\rightarrow\mathbb{R} satisfies

  • (1)

    v(y,0)v(x,0)2φ(d(x,y)2,0),x,yMv(y,0)-v(x,0)\leq 2\varphi(\frac{d(x,y)}{2},0),\ \forall x,y\in M,

  • (2)

    0φt|𝔏φ|p2p1𝔏φ,(s,t)[0,D/2]×+0\geq\varphi_{t}\geq|\mathfrak{L}\varphi|^{-\frac{p-2}{p-1}}\mathfrak{L}\varphi,\ (s,t)\in[0,D/2]\times\mathbb{R}_{+},

  • (3)

    φ(s,t)>0,(s,t)[0,D/2]×+\varphi^{\prime}(s,t)>0,\ (s,t)\in[0,D/2]\times\mathbb{R}_{+},

  • (4)

    φ(0,t)0,t>0\varphi(0,t)\geq 0,\ t>0.

Then

(3.3) v(y,t)v(x,t)2φ(d(x,y)2,t)\displaystyle v(y,t)-v(x,t)\leq 2\varphi(\frac{d(x,y)}{2},t)

for t>0t>0 and x,yMx,y\in M.

Proof.

For ε>0\varepsilon>0, let

Aε(x,y,t)=v(y,t)v(x,t)2φ(d(x,y)2,t)εet.\displaystyle A_{\varepsilon}(x,y,t)=v(y,t)-v(x,t)-2\varphi(\frac{d(x,y)}{2},t)-\varepsilon e^{t}.

To prove (3.3), it suffices to show that Aε<0A_{\varepsilon}<0 for any ε>0\varepsilon>0.

We argue by contradiction and assume that there exists (x0,y0,t0)(x_{0},y_{0},t_{0}) such that AεA_{\varepsilon} attains its maximum zero on M×M×[0,t0]M\times M\times[0,t_{0}] at (x0,y0,t0)(x_{0},y_{0},t_{0}). Clearly x0y0x_{0}\neq y_{0}, and t0>0t_{0}>0. If M\partial M\neq\emptyset, similarly as in [14, Theorem 3.1], the strictly convexity of boundary, the Neumann condition and the positivity of φ\varphi^{\prime} rule out the possibility that x0Mx_{0}\in\partial M and y0My_{0}\in\partial M.

Compactness of MM implies that there exists an arc-length minimizing geodesic γ0\gamma_{0} connecting x0x_{0} and y0y_{0} such that γ0(s0)=x0\gamma_{0}(-s_{0})=x_{0} and γ0(s0)=y0\gamma_{0}(s_{0})=y_{0} with s0=d(x0,y0)/2s_{0}=d(x_{0},y_{0})/2. We pick an orthonormal basis {ei(s)}i=12m\{e_{i}(s)\}_{i=1}^{2m} for Tx0MT_{x_{0}}M with e1=γ0(s0)e_{1}=\gamma^{\prime}_{0}(-s_{0}) and e2=Jγ0(s0)e_{2}=J\gamma_{0}^{\prime}(-s_{0}), where JJ is the complex structure. Then parallel transport along γ0\gamma_{0} produces an orthonormal basis {ei(s)}i=12m\{e_{i}(s)\}_{i=1}^{2m} for Tγ0(s)MT_{\gamma_{0}(s)}M with e1(s)=γ0(s)e_{1}(s)=\gamma_{0}^{\prime}(s) for each s[s0,s0]s\in[-s_{0},s_{0}]. Since JJ is parallel and γ0\gamma_{0} is a geodesic, we have e2(s)=Jγ0(s)e_{2}(s)=J\gamma_{0}^{\prime}(s) for each s[s0,s0]s\in[-s_{0},s_{0}].

First derivative inequality yields

(3.4) 0tAε(x0,y0,t0)=vt(y0,t0)vt(x0,t0)2φtεet0,0\leq\partial_{t}A_{\varepsilon}(x_{0},y_{0},t_{0})=v_{t}(y_{0},t_{0})-v_{t}(x_{0},t_{0})-2\varphi_{t}-\varepsilon e^{t_{0}},

and

xAε(x0,y0,t0)=yAε(x0,y0,t0)=0,\nabla_{x}A_{\varepsilon}(x_{0},y_{0},t_{0})=\nabla_{y}A_{\varepsilon}(x_{0},y_{0},t_{0})=0,

namely

(3.5) v(x0,t0)=φ(s0,t0)e1(s0),v(y0,t0)=φ(s0,t0)e1(s0).\nabla v(x_{0},t_{0})=\varphi^{\prime}(s_{0},t_{0})e_{1}(-s_{0}),\nabla v(y_{0},t_{0})=\varphi^{\prime}(s_{0},t_{0})e_{1}(s_{0}).

Recall that

(3.6) Δpv=|v|p2Δv+(p2)|v|p42v(v,v),\Delta_{p}v=|\nabla v|^{p-2}\Delta v+(p-2)|\nabla v|^{p-4}\nabla^{2}v(\nabla v,\nabla v),

then plugging equality (3.5) into (3.6) we get

(3.7) Δpv(y0,t0)Δpv(x0,t0)=(p1)|φ|p2(v11(y0,t0)v11(x0,t0))+|φ|p2i=22m(vii(y0,t0)vii(x0,t0)).\displaystyle\begin{split}&\Delta_{p}v(y_{0},t_{0})-\Delta_{p}v(x_{0},t_{0})\\ =&(p-1)|\varphi^{\prime}|^{p-2}(v_{11}(y_{0},t_{0})-v_{11}(x_{0},t_{0}))+|\varphi^{\prime}|^{p-2}\sum_{i=2}^{2m}(v_{ii}(y_{0},t_{0})-v_{ii}(x_{0},t_{0})).\end{split}

Now we use the first and second variation formulas of arc length to calculate the second derivatives in space variables. Suppose γ(r,s):[δ,δ]×[s0,s0]M\gamma(r,s):[-\delta,\delta]\times[-s_{0},s_{0}]\rightarrow M is a smooth variation of γ0\gamma_{0} such that γ(0,s)=γ0(s)\gamma(0,s)=\gamma_{0}(s), then the variation formulas give

(3.8) ddr|r=0L[γ(r,s)]=g(T,γr)|s0s0,\displaystyle\frac{\mathrm{d}}{\mathrm{d}r}\big{|}_{r=0}L[\gamma(r,s)]=g(T,\gamma_{r})\big{|}_{-s_{0}}^{s_{0}},

and

(3.9) d2dr2|r=0L[γ(r,s)]=s0s0(|(sγr)|2R(T,γr,T,γr))ds+g(T,rγr)|s0s0,\displaystyle\frac{\mathrm{d}^{2}}{\mathrm{d}r^{2}}\big{|}_{r=0}L[\gamma(r,s)]=\int_{-s_{0}}^{s_{0}}(|(\nabla_{s}\gamma_{r})^{\perp}|^{2}-R(T,\gamma_{r},T,\gamma_{r}))\ \mathrm{d}s+g(T,\nabla_{r}\gamma_{r})\big{|}_{-s_{0}}^{s_{0}},

where T=sγ(0,s)T=\partial_{s}\gamma(0,s) is the unit tangent vector to γ0\gamma_{0}.

To calculate the second derivative along e1e_{1}, we consider the variation γ(r,s):=γ0(s+rss0)\gamma(r,s):=\gamma_{0}(s+r\frac{s}{s_{0}}), and T=γ0(s)T=\gamma_{0}^{\prime}(s) and γr=γ0(s)s/s0\gamma_{r}=\gamma_{0}^{\prime}(s)s/s_{0}. Then formulas (3.8) and (3.9) give

ddr|r=0L[γ(r)]=2,andd2dr2|r=0L[γ(r)]=0.\frac{\mathrm{d}}{\mathrm{d}r}\big{|}_{r=0}L[\gamma(r)]=2,\quad\text{and}\quad\frac{\mathrm{d}^{2}}{\mathrm{d}r^{2}}\big{|}_{r=0}L[\gamma(r)]=0.

Hence the second derivative test for this variation produces

(3.10) v11(y0,t0)v11(x0,t0)2φ′′(s0,t0)0.v_{11}(y_{0},t_{0})-v_{11}(x_{0},t_{0})-2\varphi^{\prime\prime}(s_{0},t_{0})\leq 0.

To calculate the second derivative along e2e_{2}, we denote by

(3.11) cκ(t)={cos(κt),κ>0,1,κ=0,cosh(κt),κ<0,c_{\kappa}(t)=\begin{cases}\operatorname{cos}(\sqrt{\kappa}t),&\kappa>0,\\ 1,&\kappa=0,\\ \operatorname{cosh}(\sqrt{-\kappa}t),&\kappa<0,\end{cases}

and consider the variation

γ(r,s):=expγ0(s)(rη(s)e2(s)),\gamma(r,s):=\operatorname{exp}_{\gamma_{0}(s)}(r\eta(s)e_{2}(s)),

where η(s)=c4κ1(s)c4κ1(s0)\displaystyle\eta(s)=\frac{c_{4\kappa_{1}}(s)}{c_{4\kappa_{1}}(s_{0})}. Clearly T=e1(s)T=e_{1}(s) and γr=η(s)e2(s)\gamma_{r}=\eta(s)e_{2}(s), and formulas (3.8) and (3.9) imply

ddr|r=0L[γ(r)]=0\frac{\mathrm{d}}{\mathrm{d}r}\bigg{|}_{r=0}L[\gamma(r)]=0

and

d2dr2|r=0L[γ(r)]=s0s0((η)2η2R(e1,e2,e1,e2))ds.\frac{\mathrm{d}^{2}}{\mathrm{d}r^{2}}\bigg{|}_{r=0}L[\gamma(r)]=\int_{-s_{0}}^{s_{0}}\bigg{(}(\eta^{\prime})^{2}-\eta^{2}R(e_{1},e_{2},e_{1},e_{2})\bigg{)}\ \mathrm{d}s.

So this variation produces

v22(y0,t0)v22(x0,t0)φ(s0,t0)s0s0((η)2η2R(e1,e2,e1,e2))ds0.v_{22}(y_{0},t_{0})-v_{22}(x_{0},t_{0})-\varphi^{\prime}(s_{0},t_{0})\int_{-s_{0}}^{s_{0}}\bigg{(}(\eta^{\prime})^{2}-\eta^{2}R(e_{1},e_{2},e_{1},e_{2})\bigg{)}\ \mathrm{d}s\leq 0.

Using the assumption that H4κ1H\geq 4\kappa_{1} and integration by parts, we estimate that

s0s0((η)2η2R(e1,e2,e1,e2))ds\displaystyle\int_{-s_{0}}^{s_{0}}\bigg{(}(\eta^{\prime})^{2}-\eta^{2}R(e_{1},e_{2},e_{1},e_{2})\bigg{)}\ \mathrm{d}s
=\displaystyle= ηη|s0s0s0s0η2(R(e1,e2,e1,e2)4κ1)ds\displaystyle\eta^{\prime}\eta|_{-s_{0}}^{s_{0}}-\int_{-s_{0}}^{s_{0}}\eta^{2}\big{(}R(e_{1},e_{2},e_{1},e_{2})-4\kappa_{1}\big{)}\ \mathrm{d}s
\displaystyle\leq ηη|s0s0s0s0η2(R(e1,e2,e1,e2)4κ1)ds\displaystyle\eta^{\prime}\eta|_{-s_{0}}^{s_{0}}-\int_{-s_{0}}^{s_{0}}\eta^{2}\big{(}R(e_{1},e_{2},e_{1},e_{2})-4\kappa_{1}\big{)}\ \mathrm{d}s
=\displaystyle= 2T4κ1(s0).\displaystyle-2T_{4\kappa_{1}}(s_{0}).

Therefore we conclude from the above two inequalities that

(3.12) v22(y0,t0)v22(x0,t0)2T4κ1(s0)φ(s0,t0),v_{22}(y_{0},t_{0})-v_{22}(x_{0},t_{0})\leq-2T_{4\kappa_{1}}(s_{0})\varphi^{\prime}(s_{0},t_{0}),

where we used the assumption that φ(s,t)>0\varphi^{\prime}(s,t)>0.

To calculate the second derivative along eie_{i}(3i2m3\leq i\leq 2m), we consider the variation

γ(r,s):=expγ0(s)(rζ(s)ei(s)),\gamma(r,s):=\operatorname{exp}_{\gamma_{0}(s)}(r\zeta(s)e_{i}(s)),

where ζ(s)=cκ2(s)cκ2(s0)\zeta(s)=\frac{c_{\kappa_{2}}(s)}{c_{\kappa_{2}}(s_{0})}. Similarly, the second variation formula gives

vii(y0,t0)vii(x0,t0)φ(s0,t0)s0s0((ζ)2ζ2R(e1,ei,e1,ei))ds0.v_{ii}(y_{0},t_{0})-v_{ii}(x_{0},t_{0})-\varphi^{\prime}(s_{0},t_{0})\int_{-s_{0}}^{s_{0}}\bigg{(}(\zeta^{\prime})^{2}-\zeta^{2}R(e_{1},e_{i},e_{1},e_{i})\bigg{)}\ \mathrm{d}s\leq 0.

Summing over 3i2m3\leq i\leq 2m yields

i=32m(vii(y0,t0)vii(x0,t0))φ(s0,t0)i=32ms0s0((ζ)2ζ2R(e1,ei,e1,ei))ds0.\sum_{i=3}^{2m}\big{(}v_{ii}(y_{0},t_{0})-v_{ii}(x_{0},t_{0})\big{)}-\varphi^{\prime}(s_{0},t_{0})\sum_{i=3}^{2m}\int_{-s_{0}}^{s_{0}}\bigg{(}(\zeta^{\prime})^{2}-\zeta^{2}R(e_{1},e_{i},e_{1},e_{i})\bigg{)}\ \mathrm{d}s\leq 0.

Using integration by parts and assumption that Ric2(m1)κ2\operatorname{Ric}^{\perp}\geq 2(m-1)\kappa_{2}, we have

i=32ms0s0((ζ)2ζ2R(e1,ei,e1,ei))ds\displaystyle\sum_{i=3}^{2m}\int_{-s_{0}}^{s_{0}}\bigg{(}(\zeta^{\prime})^{2}-\zeta^{2}R(e_{1},e_{i},e_{1},e_{i})\bigg{)}\mathrm{d}s
=\displaystyle= 4(m1)Tκ2(s0)s0s0ζ2(2(m1)κ2+i=32mR(e1,ei,e1,ei))ds\displaystyle-4(m-1)T_{\kappa_{2}}(s_{0})-\int_{-s_{0}}^{s_{0}}\zeta^{2}\big{(}-2(m-1)\kappa_{2}+\sum_{i=3}^{2m}R(e_{1},e_{i},e_{1},e_{i})\big{)}\mathrm{d}s
\displaystyle\leq 4(m1)Tκ2(s0),\displaystyle-4(m-1)T_{\kappa_{2}}(s_{0}),

thus we get

(3.13) i=32m(vii(y0,t0)vii(x0,t0))4(m1)Tκ2(s0)φ(s0,t0).\sum_{i=3}^{2m}\big{(}v_{ii}(y_{0},t_{0})-v_{ii}(x_{0},t_{0})\big{)}\leq-4(m-1)T_{\kappa_{2}}(s_{0})\varphi^{\prime}(s_{0},t_{0}).

Inserting inequalities (3.10), (3.12) and (3.13) into (3.7), we get

(3.14) Δpv(y0,t0)Δpv(x0,t0)2𝔏φ(s0,t0).\Delta_{p}v(y_{0},t_{0})-\Delta_{p}v(x_{0},t_{0})\leq 2\mathfrak{L}\varphi(s_{0},t_{0}).

Let h(t):=|t|p2p1th(t):=|t|^{-\frac{p-2}{p-1}}t, which is odd, increasing and convex for all t>0t>0, then (3.14) implies

h(Δpv(y0,t0))h(Δpv(x0,t0)+2𝔏φ).h(\Delta_{p}v(y_{0},t_{0}))\leq h(\Delta_{p}v(x_{0},t_{0})+2\mathfrak{L}\varphi).

Applying Lemma 2.1 of [15] with t=Δpv(x0,t0)t=\Delta_{p}v(x_{0},t_{0}) and δ=𝔏φ0\delta=-\mathfrak{L}\varphi\geq 0, we get

h(Δpv(x0,t0)+2𝔏φ)h(Δpv(x0,t0))+2h(𝔏φ),h(\Delta_{p}v(x_{0},t_{0})+2\mathfrak{L}\varphi)\leq h(\Delta_{p}v(x_{0},t_{0}))+2h(\mathfrak{L}\varphi),

that is

(3.15) 12[|Δpv|p2p1Δpv|(y0,t0)|Δpv|p2p1Δpv|(x0,t0)]|𝔏φ|p2p1𝔏φ.\frac{1}{2}\big{[}|\Delta_{p}v|^{-\frac{p-2}{p-1}}\Delta_{p}v\big{|}_{(y_{0},t_{0})}-|\Delta_{p}v|^{-\frac{p-2}{p-1}}\Delta_{p}v\big{|}_{(x_{0},t_{0})}\big{]}\leq|\mathfrak{L}\varphi|^{-\frac{p-2}{p-1}}\mathfrak{L}\varphi.

Combining (3.4) and (3.15), we get

φt|𝔏φ|p2p1𝔏φε2et0<|𝔏φ|p2p1𝔏φ,\varphi_{t}\leq|\mathfrak{L}\varphi|^{-\frac{p-2}{p-1}}\mathfrak{L}\varphi-\frac{\varepsilon}{2}e^{t_{0}}<|\mathfrak{L}\varphi|^{-\frac{p-2}{p-1}}\mathfrak{L}\varphi,

which contradicts the inequality in assumption (2), so

Aε(x,y,t)<0A_{\varepsilon}(x,y,t)<0

holds true for all x,yMx,y\in M and t>0t>0. Hence we complete the proof of Theorem 3.1. \square

On the interval [0,D/2][0,D/2], we define the following corresponding one-dimensional eigenvalue problem

σ¯1=inf{0D2|ϕ|pcκ22m2c4κ1ds0D2|ϕ|pcκ22m2c4κ1ds|ϕW1,p((0,D/2))\{0},ϕ(0)=0},\bar{\sigma}_{1}=\inf\bigg{\{}\frac{\int_{0}^{\frac{D}{2}}|\phi^{\prime}|^{p}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\ \mathrm{d}s}{\int_{0}^{\frac{D}{2}}|\phi|^{p}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\ \mathrm{d}s}\ \big{|}\phi\in W^{1,p}\big{(}(0,D/2)\big{)}\backslash\{0\},\ \phi(0)=0\bigg{\}},

where cκ(t)c_{\kappa}(t) is defined by (3.11) and DD is the diameter of MM. Noting from Remark 2.1 that Dπ/4κ1D\leq\pi/\sqrt{4\kappa_{1}} if κ1>0\kappa_{1}>0, and Dπ/κ2D\leq\pi/\sqrt{\kappa_{2}} if κ2>0\kappa_{2}>0.

Lemma 3.1.
μ¯1(m,p,κ1,κ2,D)=σ¯1(m,p,κ1,κ2,D).\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D)=\bar{\sigma}_{1}(m,p,\kappa_{1},\kappa_{2},D).
Proof.

Suppose ϕ(s)\phi(s) is an eigenfunction on [0,D/2][0,D/2] corresponding to σ¯1(m,p,κ1,κ2,D)\bar{\sigma}_{1}(m,p,\kappa_{1},\kappa_{2},D), and define a test function for μ¯1(m,p,κ1,κ2,D)\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D) by

ϕ¯(s)={ϕ(s),s[0,D/2],ϕ(s),s[D/2,0],\bar{\phi}(s)=\begin{cases}\phi(s),&s\in[0,D/2],\\ -\phi(-s),&s\in[-D/2,0],\end{cases}

then we get

μ¯1(m,p,κ1,κ2,D)σ¯1(m,p,κ1,κ2,D).\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D)\leq\bar{\sigma}_{1}(m,p,\kappa_{1},\kappa_{2},D).

On the other hand, suppose ψ(s)\psi(s) is an eigenfunction corresponding to μ¯1(m,p,κ1,κ2,D)\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D), and then direct calculation gives

(|ψ|p2ψcκ22m2c4κ1)=cκ22m2c4κ1μ¯1|ψ|p2ψ.(|\psi^{\prime}|^{p-2}\psi^{\prime}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}})^{\prime}=-c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\bar{\mu}_{1}|\psi|^{p-2}\psi.

Integrating the above equation over [D/2,D/2][-D/2,D/2] yields

0=D2D2cκ22m2c4κ1|ψ|p2ψds,0=\int_{-\frac{D}{2}}^{\frac{D}{2}}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}|\psi|^{p-2}\psi\ \mathrm{d}s,

so there exists s0(D/2,D/2)s_{0}\in(-D/2,D/2) such that ψ(s0)=0\psi(s_{0})=0. Without loss of generality, we assume further that s0[0,D/2)s_{0}\in[0,D/2), and define ψ(s)=0\psi(s)=0 for s[0,s0)s\in[0,s_{0}). Then ψ\psi is a test function for σ¯1\bar{\sigma}_{1}, and we have

σ¯1(m,p,κ1,κ2,D)\displaystyle\bar{\sigma}_{1}(m,p,\kappa_{1},\kappa_{2},D)\leq 0D2|ψ|pcκ22m2c4κ1ds0D2|ψ|pcκ22m2c4κ1ds\displaystyle\frac{\int_{0}^{\frac{D}{2}}|\psi^{\prime}|^{p}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\ \mathrm{d}s}{\int_{0}^{\frac{D}{2}}|\psi|^{p}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\ \mathrm{d}s}
=\displaystyle= s0D2|ψ|pcκ22m2c4κ1dss0D2|ψ|pcκ22m2c4κ1ds\displaystyle\frac{\int_{s_{0}}^{\frac{D}{2}}|\psi^{\prime}|^{p}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\ \mathrm{d}s}{\int_{s_{0}}^{\frac{D}{2}}|\psi|^{p}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}\ \mathrm{d}s}
=\displaystyle= μ¯1(m,p,κ1,κ2,D).\displaystyle\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D).

Thus Lemma 3.1 holds. \square

Lemma 3.2.

There exists an odd eigenfunction ϕ\phi corresponding to μ¯1(m,p,κ1,κ2,D)\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D) satisfying

(3.16) (p1)|ϕ|p2ϕ′′(2(m1)Tκ2+T4κ1)|ϕ|p2ϕ=μ¯1(m,p,κ1,κ2,D)|ϕ|p2ϕ\displaystyle(p-1)|\phi^{\prime}|^{p-2}\phi^{\prime\prime}-(2(m-1)T_{\kappa_{2}}+T_{4\kappa_{1}})|\phi^{\prime}|^{p-2}\phi^{\prime}=-\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D)|\phi|^{p-2}\phi

on [0,D/2][0,D/2] with ϕ(s)>0\phi(s)>0 for s(0,D/2]s\in(0,D/2], ϕ(s)>0\phi^{\prime}(s)>0 for s[0,D/2)s\in[0,D/2), and ϕ(D/2)=0\phi^{\prime}(D/2)=0.

Proof.

From the proof of Lemma 3.1, we see that there exists an odd function ϕ\phi corresponding to μ¯1(m,p,κ1,κ2,D)\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D) with ϕ(0)=0\phi(0)=0 and ϕ(D/2)=0\phi^{\prime}(D/2)=0, which does not change sign in (0,D/2](0,D/2]. Since equation (3.16) is equivalent to

(|ϕ|p2ϕcκ22m2c4κ1)=μ¯1cκ22m2c4κ1|ϕ|p2ϕ(|\phi^{\prime}|^{p-2}\phi^{\prime}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}})^{\prime}=-\bar{\mu}_{1}c_{\kappa_{2}}^{2m-2}c_{4\kappa_{1}}|\phi|^{p-2}\phi

and ϕ(D/2)=0\phi^{\prime}(D/2)=0, then ϕ(s)>0\phi^{\prime}(s)>0 on [0,D/2)[0,D/2). \square

Now we turn to prove Theorem 1.2.

Proof of Theorem 1.2.

For any D1>DD_{1}>D, let μ¯1=μ¯1(m,p,κ1,κ2,D1)\bar{\mu}_{1}=\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D_{1}) be the first nonzero Neumann eigenvalue of the eigenvalue problem (1.5). By Lemma 3.2, there exists an odd eigenfunction ϕ(s)\phi(s) corresponding to μ¯1\bar{\mu}_{1} such that ϕ(s)>0\phi(s)>0 on (0,D1/2](0,D_{1}/2], ϕ(s)>0\phi^{\prime}(s)>0 on [0,D1/2)[0,D_{1}/2) and ϕ(D1/2)=0\phi^{\prime}(D_{1}/2)=0. Let u(x)u(x) be an eigenfucntion corresponding to μ1,p(M)\mu_{1,p}(M), and it is well known that u(x)C1,α(M)u(x)\in C^{1,\alpha}(M) for some α(0,1)\alpha\in(0,1), then there exists C1>0C_{1}>0 such that

(3.17) |u(y)u(x)|C1d(x,y)|u(y)-u(x)|\leq C_{1}d(x,y)

for x,yMx,y\in M. Noting that φ\varphi^{\prime} is positive on [0,D/2][0,D/2], then from equation (3.16) we deduce that φ\varphi is smooth on [0,D/2][0,D/2], and for all x,yMx,y\in M it holds

(3.18) ϕ(d(x,y)2)C2d(x,y)\phi(\frac{d(x,y)}{2})\geq C_{2}d(x,y)

for some C2>0C_{2}>0. By (3.17) and (3.18), we can choose a C>0C>0 such that

|u(y)u(x)|2Cϕ(d(x,y)2)|u(y)-u(x)|\leq 2C\phi(\frac{d(x,y)}{2})

for all x,yMx,y\in M.

Let

v(x,t)=exp(μ1,p1p1t)u(x),\displaystyle v(x,t)=\exp(-\mu_{1,p}^{\frac{1}{p-1}}t)u(x),

and

φ(s,t)=Cexp(μ¯11p1t)ϕ(s).\displaystyle\varphi(s,t)=C\exp(-\bar{\mu}_{1}^{\frac{1}{p-1}}t)\phi(s).

It is easy to verify that vv and φ\varphi satisfy the conditions in Theorem 3.1, and then it holds

v(y,t)v(x,t)2φ(d(x,y)2,t)v(y,t)-v(x,t)\leq 2\varphi(\frac{d(x,y)}{2},t)

for all t>0t>0, namely

exp(μ1,p1p1t)(u(y)u(x))2Cexp(μ¯11p1t)ϕ(d(x,y)2).\exp(-\mu_{1,p}^{\frac{1}{p-1}}t)\big{(}u(y)-u(x)\big{)}\leq 2C\exp(-\bar{\mu}_{1}^{\frac{1}{p-1}}t)\phi(\frac{d(x,y)}{2}).

As tt\rightarrow\infty, the above inequality gives

μ1,pμ¯1(m,p,κ1,κ2,D1),\mu_{1,p}\geq\bar{\mu}_{1}(m,p,\kappa_{1},\kappa_{2},D_{1}),

then Theorem 1.2 follows by letting DD1D\rightarrow D_{1}. \square

4. The First Dirichlet Eigenvalue

In this section, we give the proof of Theorem 1.4. The key tools are a comparison theorem for distance to the boundary and Barta’s inequality. Let MM be a compact manifold with smooth boundary M\partial M, and define the distance function to the boundary of MM by

d(x,M)=inf{d(x,y):yM}.d(x,\partial M)=\inf\Big{\{}d(x,y):y\in\partial M\Big{\}}.

By choosing α=(p1)|ψ|p2\alpha=(p-1)|\nabla\psi|^{p-2} and β=|ψ|p2\beta=|\nabla\psi|^{p-2} in Theorem 6.1 of [14], we have the following lemma.

Lemma 4.1.

Let (Mm,g,J)(M^{m},g,J) be a compact Kähler manifold with smooth boundary M\partial M. Suppose that H4κ1H\geq 4\kappa_{1} and Ric2(m1)κ2\operatorname{Ric}^{\perp}\geq 2(m-1)\kappa_{2} for some κ1,κ2\kappa_{1},\kappa_{2}\in\mathbb{R}, and the second fundamental form on M\partial M is bounded from below by Λ\Lambda\in\mathbb{R}. Let p(1,)p\in(1,\infty), and assume φ\varphi is a smooth function on [0,R][0,R] satisfying φ0\varphi^{\prime}\geq 0. Then for any smooth function ψ\psi satisfying

ψ(x)φ(d(x,M))forxM,andψ(x0)=φ(d(x0,M)),\psi(x)\leq\varphi(d(x,\partial M))\quad\text{for}\quad x\in M,\quad\text{and}\quad\psi(x_{0})=\varphi(d(x_{0},\partial M)),

it holds

Δpψ(x0)𝔏¯φ(d(x0,M)),\Delta_{p}\psi(x_{0})\leq\bar{\mathfrak{L}}\varphi(d(x_{0},\partial M)),

where one-dimensional operator 𝔏¯\bar{\mathfrak{L}} is defined by

(4.1) 𝔏¯φ=(p1)|φ|p2φ′′(2(m1)Tκ2,Λ+T4κ1,Λ)|φ|p2φ\displaystyle\bar{\mathfrak{L}}\varphi=(p-1)|\varphi^{\prime}|^{p-2}\varphi^{\prime\prime}-\big{(}2(m-1)T_{\kappa_{2},\Lambda}+T_{4\kappa_{1},\Lambda}\big{)}|\varphi^{\prime}|^{p-2}\varphi^{\prime}

for all φC2([0,R])\varphi\in C^{2}([0,R]), and Tκ,ΛT_{\kappa,\Lambda} is defined by (1.8).

Consider the following one-dimensional eigenvalue problem on [0,R][0,R]

(4.2) {𝔏¯φ=λ|φ|p2φ,φ(0)=0,φ(R)=0,\left\{\begin{aligned} &\bar{\mathfrak{L}}\varphi=-\lambda|\varphi|^{p-2}\varphi,\\ &\varphi(0)=0,\ \varphi^{\prime}(R)=0,\end{aligned}\right.

and denote by λ¯1(m,p,κ1,κ2,D)\bar{\lambda}_{1}(m,p,\kappa_{1},\kappa_{2},D), written as λ¯1\bar{\lambda}_{1} for short, the first eigenvalue of problem (4.2). Then it is easy to see that λ¯1\bar{\lambda}_{1} can be characterized by

λ¯1=inf{0R|ϕ|pCκ2,Λ2m2C4κ1,Λds0R|ϕ|pCκ2,Λ2m2C4κ1,Λds|ϕW1,p((0,R))\{0},ϕ(0)=0},\bar{\lambda}_{1}=\inf\bigg{\{}\frac{\int_{0}^{R}|\phi^{\prime}|^{p}C_{\kappa_{2},\Lambda}^{2m-2}C_{4\kappa_{1},\Lambda}\ \mathrm{d}s}{\int_{0}^{R}|\phi|^{p}C_{\kappa_{2},\Lambda}^{2m-2}C_{4\kappa_{1},\Lambda}\ \mathrm{d}s}\ \big{|}\phi\in W^{1,p}\big{(}(0,R)\big{)}\backslash\{0\},\ \phi(0)=0\bigg{\}},

where Cκ,ΛC_{\kappa,\Lambda} is defined by (1.7).

Proof of Theorem 1.4.

Let φ\varphi be an eigenfunction corresponding to λ¯1\bar{\lambda}_{1}, then

𝔏¯φ=λ¯1|φ|p2φ\bar{\mathfrak{L}}\varphi=-\bar{\lambda}_{1}|\varphi|^{p-2}\varphi

with φ(0)=0\varphi(0)=0 and φ(R)=0\varphi^{\prime}(R)=0. Similarly as in the proof of Lemma 3.2, we can choose φ\varphi such that φ(s)>0\varphi(s)>0 on (0,R](0,R], and φ(s)>0\varphi^{\prime}(s)>0 on [0,R)[0,R). Define a trial function for λ1,p(M)\lambda_{1,p}(M) by

v(x):=φ(d(x,M)),v(x):=\varphi(d(x,\partial M)),

then Lemma 4.1 gives

(4.3) Δpv(x)𝔏¯φ(d(x,M))\displaystyle\Delta_{p}v(x)\leq\bar{\mathfrak{L}}\varphi(d(x,\partial M))

away from the cut locus of M\partial M, and thus globally in the distributional sense, see [13, Lemma 5.2]. Recall that φ\varphi is an eigenfunction with respect to λ¯1\bar{\lambda}_{1}, namely

(4.4) 𝔏¯φ(d(x,M))=λ¯1|v(x)|p2v(x),\displaystyle\bar{\mathfrak{L}}\varphi(d(x,\partial M))=-\bar{\lambda}_{1}|v(x)|^{p-2}v(x),

thus we conclude from (4.3) and (4.4) that

Δpvλ¯1|v|p2v,xM.\displaystyle\Delta_{p}v\leq-\bar{\lambda}_{1}|v|^{p-2}v,\quad x\in M.

By the definition of v(x)v(x), we see v(x)>0v(x)>0 for xMx\in M, and v(x)=0v(x)=0 for xMx\in\partial M. Then using Barta’s inequality (cf. [13, Theorem 3.1]), we have

λ1,p(M)λ¯1,\lambda_{1,p}(M)\geq\bar{\lambda}_{1},

proving Theorem 1.4. \square

Lastly, we emphasize that when κ1=κ2=κ\kappa_{1}=\kappa_{2}=\kappa, a standard argument (see for instance [13, Section 6]) indicates that when MM is a (κ,Λ)(\kappa,\Lambda)-model space in the complex space forms, the first Dirichlet eigenfunction of the pp-Laplacian can be written in the form u=φd(x,M)u=\varphi\circ d(x,\partial M), and φ\varphi is the first eigenfunction of one-dimensional problem (4.2), which gives the equality case of Theorem 1.2. Therefore estimate (1.9) is sharp when κ1=κ2\kappa_{1}=\kappa_{2}.

References

  • [1] Ben Andrews. Moduli of continuity, isoperimetric profiles, and multi-point estimates in geometric heat equations. In Surveys in differential geometry 2014. Regularity and evolution of nonlinear equations, volume 19 of Surv. Differ. Geom., pages 1–47. Int. Press, Somerville, MA, 2015.
  • [2] Ben Andrews and Julie Clutterbuck. Time-interior gradient estimates for quasilinear parabolic equations. Indiana Univ. Math. J., 58(1):351–380, 2009.
  • [3] Ben Andrews and Julie Clutterbuck. Proof of the fundamental gap conjecture. J. Amer. Math. Soc., 24(3):899–916, 2011.
  • [4] Ben Andrews and Julie Clutterbuck. Sharp modulus of continuity for parabolic equations on manifolds and lower bounds for the first eigenvalue. Anal. PDE, 6(5):1013–1024, 2013.
  • [5] Casey Blacker and Shoo Seto. First eigenvalue of the pp-Laplacian on Kähler manifolds. Proc. Amer. Math. Soc., 147(5):2197–2206, 2019.
  • [6] Mu Fa Chen and Feng Yu Wang. Application of coupling method to the first eigenvalue on manifold. Sci. China Ser. A, 37(1):1–14, 1994.
  • [7] Chenxu He, Guofang Wei, and Qi S. Zhang. Fundamental gap of convex domains in the spheres. Amer. J. Math., 142(4):1161–1191, 2020.
  • [8] Atsushi Kasue. On a lower bound for the first eigenvalue of the Laplace operator on a Riemannian manifold. Ann. Sci. École Norm. Sup. (4), 17(1):31–44, 1984.
  • [9] Pawel Kröger. On the spectral gap for compact manifolds. J. Differential Geom., 36(2):315–330, 1992.
  • [10] Peter Li. A lower bound for the first eigenvalue of the Laplacian on a compact manifold. Indiana Univ. Math. J., 28(6):1013–1019, 1979.
  • [11] Peter Li and Jiaping Wang. Comparison theorem for Kähler manifolds and positivity of spectrum. J. Differential Geom., 69(1):43–74, 2005.
  • [12] 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.
  • [13] Xiaolong Li and Kui Wang. First Robin eigenvalue of the pp-Laplacian on Riemannian manifolds. Math. Z., 298(3-4):1033–1047, 2021.
  • [14] Xiaolong Li and Kui Wang. Lower bounds for the first eigenvalue of the Laplacian on Kähler manifolds. Trans. Amer. Math. Soc., 374(11):8081–8099, 2021.
  • [15] Xiaolong Li and Kui Wang. Sharp lower bound for the first eigenvalue of the weighted pp-Laplacian I. J. Geom. Anal., 31(8):8686–8708, 2021.
  • [16] Xiaolong Li and Kui Wang. Sharp lower bound for the first eigenvalue of the weighted pp-laplacian II. Math. Res. Lett., 28(5):1459–1479, 2021.
  • [17] André Lichnerowicz. Géométrie des groupes de transformations. Travaux et Recherches Mathématiques, III. Dunod, Paris, 1958.
  • [18] Ovidiu Munteanu. A sharp estimate for the bottom of the spectrum of the Laplacian on Kähler manifolds. J. Differential Geom., 83(1):163–187, 2009.
  • [19] 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.
  • [20] Lei Ni. Estimates on the modulus of expansion for vector fields solving nonlinear equations. J. Math. Pures Appl. (9), 99(1):1–16, 2013.
  • [21] Lei Ni and Fangyang Zheng. Comparison and vanishing theorems for Kähler manifolds. Calc. Var. Partial Differential Equations, 57(6):Paper No. 151, 31, 2018.
  • [22] Benjamin Rutkowski and Shoo Seto. Explicit lower bound of the first eigenvalue of the laplacian on kähler manifolds. arXiv:2207.10966, 2022.
  • [23] Shoo Seto, Lili Wang, and Guofang Wei. Sharp fundamental gap estimate on convex domains of sphere. J. Differential Geom., 112(2):347–389, 2019.
  • [24] Yôtarô Tsukamoto. On Kählerian manifolds with positive holomorphic sectional curvature. Proc. Japan Acad., 33:333–335, 1957.
  • [25] Daniele Valtorta. Sharp estimate on the first eigenvalue of the pp-Laplacian. Nonlinear Anal., 75(13):4974–4994, 2012.
  • [26] Yuntao Zhang and Kui Wang. An alternative proof of lower bounds for the first eigenvalue on manifolds. Math. Nachr., 290(16):2708–2713, 2017.
  • [27] 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.