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

The second Robin eigenvalue in non-compact rank-1 symmetric spaces

Xiaolong Li Department of Mathematics, Statistics and Physics, Wichita State University, Wichita, KS 67260, USA [email protected] Kui Wang School of Mathematical Sciences, Soochow University, Suzhou, 215006, China [email protected]  and  Haotian Wu School of Mathematics and Statistics, The University of Sydney, NSW 2006, Australia [email protected]
Abstract.

In this paper, we prove a quantitative spectral inequality for the second Robin eigenvalue in non-compact rank-1 symmetric spaces. In particular, this shows that for bounded domains in non-compact rank-1 symmetric spaces, the geodesic ball maximises the second Robin eigenvalue among domains of the same volume, with negative Robin parameter in the regime connecting the first nontrivial Neumann and Steklov eigenvalues. This result generalises the work of Freitas and Laugesen in the Euclidean setting [16] as well as our previous work in the hyperbolic space [19].

Key words and phrases:
Second Robin Eigenvalue, Eigenvalue Comparison, Rank-1 Symmetric Space
2010 Mathematics Subject Classification:
35P15, 49R05, 58C40, 58J50

1. Introduction

The Robin eigenvalue problem is to solve

{Δu=λu in Ω,uν+αu=0 on Ω,\displaystyle\begin{cases}-\Delta u=\lambda u&\text{ in }\Omega,\\ \frac{\partial u}{\partial\nu}+\alpha u=0&\text{ on }\partial\Omega,\end{cases} (1.1)

where Ω\Omega is a bounded domain with Lipschitz boundary, Δ\Delta denotes the Laplace-Beltrami operator, ν\nu denotes the outward unit normal to Ω\partial\Omega, and α\alpha\in\mathbb{R} is the Robin parameter. The eigenvalues, denoted by λk,α(Ω)\lambda_{k,\alpha}(\Omega) for k=1,2,k=1,2,\cdots, are increasing and continuous in α\alpha, and for each α\alpha satisfy

λ1,α(Ω)λ2,α(Ω)λ3,α(Ω),\displaystyle\lambda_{1,\alpha}(\Omega)\leq\lambda_{2,\alpha}(\Omega)\leq\lambda_{3,\alpha}(\Omega)\leq\cdots\rightarrow\infty,

where each eigenvalue is repeated according to its multiplicity. The first eigenvalue is simple if Ω\Omega is connected; the first eigenfunction is positive. The Robin eigenvalue problem generates a global picture of the spectrum of the Laplace operator. Indeed, the Neumann (α=0)(\alpha=0), the Steklov (λ=0\lambda=0 and α-\alpha equals the Steklov eigenvalue) and the Dirichlet (α+)\alpha\to+\infty) eigenvalue problems are all special cases of the Robin eigenvalue problem. Hence, existing results on the Dirichlet, Neumann or Steklov eigenvalues naturally motivate the investigation on the Robin eigenvalues.

In Euclidean space, the classical Faber-Krahn inequality asserts that the ball uniquely minimizes the first Dirichlet eigenvalue among bounded domains with the same volume. When α>0\alpha>0, the ball is also the unique minimizer of the first Robin eigenvalue λ1,α(Ω)\lambda_{1,\alpha}(\Omega) among domains of the same volume in n\mathbb{R}^{n}, as was shown in dimension two by Bossel [5] in 1986 and extended to all dimensions n2n\geq 2 by Daners [12] in 2006. An alternative approach via the calculus of variations was found by Bucur and Giacomini [9, 10] later. For negative values of α\alpha, it was conjectured by Bareket [3] in 1977 that the ball would be the maximizer among domains in n\mathbb{R}^{n} with the same volume. However, in 2015, Freitas and Krejčiřik [14] disproved Bareket’s conjecture by showing that the ball is not a maximizer for sufficiently negative values of α\alpha. In the same paper, the authors showed that in dimension two, the disk uniquely maximizes λ1,α(Ω)\lambda_{1,\alpha}(\Omega) for α<0\alpha<0 with |α||\alpha| sufficiently small, and conjectured that the maximizer still has radial symmetry whenever α<0\alpha<0 and should switch from a ball to a shell at some critical value of α\alpha.

Let us turn to the shape optimization problem for the second Robin eigenvalue λ2,α(Ω)\lambda_{2,\alpha}(\Omega). Suppose for the moment that Ωn\Omega\subset\mathbb{R}^{n}. When α>0\alpha>0, both the second Dirichlet eigenvalue and the second Robin eigenvalue are uniquely minimized by the disjoint union of two equal balls among bounded Lipschitz domains of the same volume. This was proved by Kennedy [18]. When α=0\alpha=0, we have λ2,0(Ω)=μ1(Ω)\lambda_{2,0}(\Omega)=\mu_{1}(\Omega), the first nonzero Neumann eigenvalue, for which the classical Szegö-Weinberger inequality [20, 21] states that among domains with the same volume, the ball uniquely maximizes μ1(Ω)\mu_{1}(\Omega), see [7] for stability version as well. When α<0\alpha<0, it is expected, cf. [17, Problem 4.41], that λ2,α(Ω)\lambda_{2,\alpha}(\Omega) should be maximal on the ball for a range of Robin parameters. This expectation has recently been confirmed by Freitas and Laugesen via Szegö way [15] and Weineberger way [16] respectively. Precisely the following theorem holds true.

Theorem 1.1 (Theorem A of [16]).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded Lipschitz domain, n2n\geq 2 and BB be a round ball of the same volume as Ω\Omega. If α[n+1nR,0]\alpha\in[-\frac{n+1}{n}R,0], then

λ2,α(Ω)λ2,α(B),\displaystyle\lambda_{2,\alpha}(\Omega)\leq\lambda_{2,\alpha}(B), (1.2)

where RR is the radius of BB. Equality holds if and only if Ω\Omega is a round ball.

Inequality (1.2) asserts that the ball uniquely maximizes λ2,α(Ω)\lambda_{2,\alpha}(\Omega) among domains of the same volume provided that α\alpha lies in a regime connecting the first nonzero Neumann eigenvalue μ1\mu_{1} and the first nonzero Steklov eigenvalue σ1\sigma_{1}, namely α[n+1nR1,0]\alpha\in[-\frac{n+1}{n}R^{-1},0]. Taking α=0\alpha=0 and α=1/R\alpha=-1/R recovers the Szegö-Weinberger inequality [22] for μ1(Ω)\mu_{1}(\Omega) and the Brock-Weinstock inequality [8] for σ1(Ω)\sigma_{1}(\Omega) respectively, and both classical inequalities assert that the ball is the unique maximizer among domains with the same volume in Euclidean space. We mention that the stability version of the Brock-Weinstock inequality has been proved in [6].

It is known that the Szegö-Weinberger inequality holds for domains in the hemisphere, in the hyperbolic space [2] and in rank-1 symmetric spaces (ROSS) [1], asserting that the geodesic ball maximises the first nonzero Neumann eigenvalues among domains of the same volume. Binoy and Santhanam proved [4] the Brock-Weinstock inequality in hyperbolic space and non-compact ROSS, and Castillon and Ruffini proved [11] a stability version of the Brock-Weinstock inequality in non-compact ROSS and that the Brock-Weinstock inequality does not hold on sphere in general. Recently, in their work [19], the authors have extended the result of Freitas and Laugesen [16] to bounded domains in real hyperbolic spaces, and proved that in complete simply connected nonpositively curved space forms, geodesic balls uniquely maximize the second Robin eigenvalue among domains with the same volume, with negative Robin parameter in the regime connecting the first nontrivial Neumann and Steklov eigenvalues.

In this paper, we continue our study of the optimization problem for the second Robin eigenvalue for some negative Robin parameter [19], and prove the following quantitative Freitas-Laugesen inequality in non-compact ROSS.

Theorem 1.2.

Let Mm=𝕂HnM^{m}=\mathbb{K}H^{n} be a non-compact ROSS of real dimension m=knm=kn (cf. Section 2). For any v>0v>0, there exists a constant C=C(m,k,v)C=C(m,k,v) such that for any domain ΩM\Omega\subset M with |Ω|=v|\Omega|=v, any geodesic ball BB with |B|=v|B|=v, and any α[σ1(B),0]\alpha\in[-\sigma_{1}(B),0], we have

λ2,α(Ω)+C𝒜2(Ω)λ2,α(B),\displaystyle\lambda_{2,\alpha}(\Omega)+C\mathcal{A}^{2}(\Omega)\leq\lambda_{2,\alpha}(B), (1.3)

where 𝒜(Ω)\mathcal{A}(\Omega) is the Fraenkel asymmetry of Ω\Omega defined by

𝒜(Ω)=inf{|ΩB|+|BΩ||B|:Bis any geodesic ball,|Ω|=|B|}.\mathcal{A}(\Omega)=\inf\left\{\frac{|\Omega\setminus B|+|B\setminus\Omega|}{|B|}~{}:~{}B~{}\text{is any geodesic ball},~{}|\Omega|=|B|\right\}.

Equality holds if and only if Ω\Omega is a geodesic ball.

By taking α=0\alpha=0, Theorem 1.2 yields that geodesic balls uniquely maximize the first nonzero Neumann eigenvalue among domains of the same volume in noncomapct ROSS, recovering Theorem 2 of [1] proved by Aithal and Santhanam. By taking α=σ1(B)\alpha=-\sigma_{1}(B), Theorem 1.2 implies that geodesic balls uniquely maximize the first nonzero Steklov eigenvalue among domains of the same volume in noncomapct ROSS, which recovers the result of Binoy and Santhanam in [4], and the result of Castillon and Ruffini in [11]. Noticing from [11] that the Brock-Weinstock inequality does not hold on sphere in general, so we mention here that Theorem 1.2 does not hold in compact space forms or compact ROSS.

This paper is organized as follows. In Section 2, we collect some known facts on non-comapct ROSS. In Section 3, we prove some properties of second Robin eigenvalues and eigenfuctions for geodesic balls in noncomapct ROSS. In Section 4, we prove a monotonicity lemma, which is useful in the proof of Theorem 1.2. Section 5 is devoted to the proof of Theorem 1.2.

Acknowledgements

X. Li is partially supported by a start-up grant at Wichita State University; K. Wang is partially supported by NSFC No.11601359; H. Wu is partially supported by ARC Grant DE180101348.

2. Geometry of non-compact rank-1 symmetric spaces

Let 𝕂\mathbb{K} denote one of the following: the field 𝐑\mathbf{R} of real numbers, the field 𝐂\mathbf{C} of complex numbers, and the algebra 𝐇\mathbf{H} of quaternions or or the algebra 𝐂𝐚\mathbf{Ca} of octonions. Let Mm=𝕂HnM^{m}=\mathbb{K}H^{n}, a non-compact ROSS of dimension m=knm=kn, where k=dim(𝕂)k=\operatorname{dim}_{\mathbb{R}}(\mathbb{K}). It is known that MM carries k1k-1 orthogonal complex structures J1,,Jk1J_{1},\cdots,J_{k-1}, and for any unit vector X,YTxMX,Y\in T_{x}M with YY orthogonal to X,J1X,,Jk1XX,J_{1}X,\cdots,J_{k-1}X we have

Rm(X,JiX)X=4JiX, and Rm(X,Y)X=Y,\displaystyle\operatorname{Rm}(X,J_{i}X)X=-4J_{i}X,\text{\quad and \quad}\operatorname{Rm}(X,Y)X=-Y, (2.1)

where 1ik11\leq i\leq k-1 and Rm\operatorname{Rm} is the Riemannian curvature operator.

Now we collect some known facts about geodesic polar coordinates, see for example [1, Sect. 3]. In geodesic polar coordinates centered at a point pMp\in M, the Riemannian density function J(r)J(r) is given by

J(r)=sinhm1rcoshk1r,J(r)=\sinh^{m-1}r\cosh^{k-1}r,

where rr is the distance function to pp. The Laplace operator ΔM\Delta_{M} is given by

ΔM=2r2+H(r)r+ΔSr,\Delta_{M}=\frac{\partial^{2}}{\partial r^{2}}+H(r)\frac{\partial}{\partial r}+\Delta_{S_{r}},

where

H(r)=(logJ(r))=(m1)cothr+(k1)tanhr,H(r)=(\log J(r))^{\prime}=(m-1)\coth r+(k-1)\tanh r,

the mean curvature of the distance sphere Sr:={xM:dist(p,x)=r}S_{r}:=\{x\in M:\operatorname{dist}(p,x)=r\}, and ΔSr\Delta_{S_{r}} denotes the Laplacian of SrS_{r}. Moreover, the first non-zero eigenvalue of ΔSr\Delta_{S_{r}} is

λ1(Sr)=m1sinh2rk1cosh2r=H(r),\displaystyle\lambda_{1}(S_{r})=\frac{m-1}{\sinh^{2}r}-\frac{k-1}{\cosh^{2}r}=-H^{\prime}(r), (2.2)

and the associated eigenfunctions are the linear coordinate functions restricted to 𝕊m1\mathbb{S}^{m-1}, denoted by ψi(θ)\psi_{i}(\theta) (1im1\leq i\leq m), satisfying

i=1m|Srψi|2=λ1(Sr)=H(r),\displaystyle\sum_{i=1}^{m}|\nabla^{S_{r}}\psi_{i}|^{2}=\lambda_{1}(S_{r})=-H^{\prime}(r), (2.3)

see Lemma 4.11 of [11] for details.

3. Robin eigenvalues for geodesic balls

The theory of self-adjoint operators yields variational characterizations of Laplacian eigenvalues. In particular, the first two Robin eigenvalues are characterized by

λ1,α(Ω)=inf{Ω|u|2𝑑μg+αΩu2𝑑AgΩu2𝑑μg:uW1,2(Ω){0}}\displaystyle\lambda_{1,\alpha}(\Omega)=\inf\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,d\mu_{g}+\alpha\int_{\partial\Omega}u^{2}\,dA_{g}}{\int_{\Omega}u^{2}\,d\mu_{g}}:u\in W^{1,2}(\Omega)\setminus\{0\}\right\} (3.1)

and

λ2,α(Ω)=inf{Ω|u|2𝑑μg+αΩu2𝑑AgΩu2𝑑μg:uW1,2(Ω){0},Ωuu1𝑑μg=0},\displaystyle\lambda_{2,\alpha}(\Omega)=\inf\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,d\mu_{g}+\alpha\int_{\partial\Omega}u^{2}\,dA_{g}}{\int_{\Omega}u^{2}\,d\mu_{g}}:u\in W^{1,2}(\Omega)\setminus\{0\},\;\int_{\Omega}uu_{1}\,d\mu_{g}=0\right\}, (3.2)

where u1u_{1} is the first eigenfunction associated with λ1,α(Ω)\lambda_{1,\alpha}(\Omega), and the first nonzero Steklov eigenvalue σ1(Ω)\sigma_{1}(\Omega) is characterized variationally by

σ1(Ω)=inf{Ω|u|2𝑑μgΩu2𝑑Ag:uW1,2(Ω){0},Ωu𝑑Ag=0},\displaystyle\sigma_{1}(\Omega)=\inf\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,d\mu_{g}}{\int_{\partial\Omega}u^{2}\,dA_{g}}:u\in W^{1,2}(\Omega)\setminus\{0\},\;\int_{\partial\Omega}u\,dA_{g}=0\right\}, (3.3)

where dμgd\mu_{g} is the Riemnnian measure induced by the metric gg and dAgdA_{g} is the induced measure on Ω\partial\Omega.

Let MM be a non-compact ROSS and BRMB_{R}\subset M be a geodesic ball of radius RR, and we consider the following Robin eigenvalue problem

{Δu(x)=λu(x),xBR,uν+αu=0,xBR.\displaystyle\begin{cases}-\Delta u(x)=\lambda u(x),&x\in B_{R},\\ \frac{\partial u}{\partial\nu}+\alpha u=0,&x\in\partial B_{R}.\end{cases} (3.4)

Since λ1,α\lambda_{1,\alpha} is simple and BRB_{R} is rotational symmetric, then the first eigenfunction is radial, hence λ1,α(BR)\lambda_{1,\alpha}(B_{R}) is the first eigenvalue of

f′′(r)H(r)f(r)=τf(r),0<r<R,\displaystyle-f^{\prime\prime}(r)-H(r)f^{\prime}(r)=\tau f(r),\quad 0<r<R, (3.5)

with f(0)=0f^{\prime}(0)=0 and f(R)+αf(R)=0f^{\prime}(R)+\alpha f(R)=0. By using the separation of variables technique, we see that the second eigenvalue λ2,α(BR)\lambda_{2,\alpha}(B_{R}) of Problem (3.4) is either the second eigenvalue of (3.5) or the first eigenvalue of

g′′(r)H(r)g(r)+λ1(Sr)g(r)=μg(r),0<r<R,\displaystyle-g^{\prime\prime}(r)-H(r)g^{\prime}(r)+\lambda_{1}(S_{r})g(r)=\mu g(r),\quad 0<r<R, (3.6)

with g(0)=0g(0)=0 and g(R)+αg(R)=0g^{\prime}(R)+\alpha g(R)=0, where λ1(Sr)\lambda_{1}(S_{r}) is given by (2.2). Moreover it’s easily seen that the first eigenvalue of (3.6) is characterized variationally by

μ1=infgW1,2((0,R)){0R(g(r)2+λ1(Sr)g2)J(r)𝑑r+αg2(R)J(R)0Rg2J(r)𝑑r:g(0)=0},\displaystyle\mu_{1}=\inf_{g\in W^{1,2}((0,R))}\left\{\frac{\int_{0}^{R}\left(g^{\prime}(r)^{2}+\lambda_{1}(S_{r})g^{2}\right)J(r)\,dr+\alpha g^{2}(R)J(R)}{\int_{0}^{R}g^{2}J(r)\,dr}:g(0)=0\right\}, (3.7)

and the associated eigenfunction g(r)g(r) can be so chosen that g(r)>0g(r)>0 for r(0,R]r\in(0,R] and g(0)=1g^{\prime}(0)=1.

We prove the following properties of the second eigenvalue and eigenfunctions of (3.4) for negative Robin parameter.

Proposition 3.1.

Suppose α0\alpha\leq 0, then the second Robin eigenfunctions of (3.4) are given by

ui(x)=g(r)ψi(θ),i=1,2,,m,\displaystyle u_{i}(x)=g(r)\psi_{i}(\theta),\quad i=1,2,\cdots,m,

where ψi(θ)\psi_{i}(\theta)’s are the linear coordinate functions restricted to 𝕊m1\mathbb{S}^{m-1}, and g(r):[0,R][0,)g(r):[0,R]\rightarrow[0,\infty) solves

g′′(r)+H(r)g(r)+(λ2,α(BR)λ1(Sr))g(r)=0g^{\prime\prime}(r)+H(r)g^{\prime}(r)+\left(\lambda_{2,\alpha}(B_{R})-\lambda_{1}(S_{r})\right)g(r)=0 (3.8)

with boundary condition g(0)=0g(0)=0 and g(R)=αg(R)g^{\prime}(R)=-\alpha g(R).

Proof.

Let μ1\mu_{1} and τ2\tau_{2} be the first nonzero eigenvalue of (3.6) and the second eigenvalue of (3.5) respectively. Then it suffices to show μ1<τ2.\mu_{1}<\tau_{2}. Let gg and ff be eigenfunctions associated to μ1\mu_{1} and τ2\tau_{2} resp., and set

h(r):=0rg(t)𝑑t1μ1g(0),h(r):=\int_{0}^{r}g(t)\,dt-\frac{1}{\mu_{1}}g^{\prime}(0),

then we deduce from (3.6) that

h′′(r)H(r)h(r)=μ1h(r),\displaystyle-h^{\prime\prime}(r)-H(r)h^{\prime}(r)=\mu_{1}h(r), (3.9)

for 0<rR0<r\leq R, particularly

μ1h(R)=g(R)H(R)g(R)=(αH(R))g(R)<0,\mu_{1}h(R)=-g^{\prime}(R)-H(R)g(R)=\big{(}\alpha-H(R)\big{)}g(R)<0,

where in the inequality we used the assumption α<0\alpha<0. Since h(r)=g(r)>0h^{\prime}(r)=g(r)>0 in (0,R)(0,R), then hh is negative in (0,R](0,R].

Using (3.5) and (3.9), we calculate

(J(fhfh))=(τ2μ1)Jfh.\displaystyle\Big{(}J\big{(}fh^{\prime}-f^{\prime}h\big{)}\Big{)}^{\prime}=(\tau_{2}-\mu_{1})Jfh. (3.10)

Recall that ff is an eigenfuction corresponding to the second eigenvalue of (3.5), it must change sign in (0,R)(0,R). Without loss of generality, we assume f(r)f(r) is positive in (0,a)(0,a) for some a<Ra<R and f(a)=0f(a)=0. Integrating (3.10) over (0,a)(0,a) yields

(τ2μ1)0aJ(t)f(t)h(t)𝑑t=J(r)(f(r)h(r)f(r)h(r))|0a=J(a)f(a)h(a).\displaystyle\begin{split}&(\tau_{2}-\mu_{1})\int_{0}^{a}J(t)f(t)h(t)\,dt\\ =&J(r)\Big{(}f(r)h^{\prime}(r)-f^{\prime}(r)h(r)\Big{)}\Big{|}_{0}^{a}\\ =&-J(a)f^{\prime}(a)h(a).\end{split} (3.11)

Since f(a)<0f^{\prime}(a)<0, and f(r)>0f(r)>0 in (0,a)(0,a), and h(r)<0h(r)<0, we then get from (3.11) that μ1<τ2,\mu_{1}<\tau_{2}, proving the proposition. \square

Proposition 3.2.

Let σ1(BR)\sigma_{1}(B_{R}) be the first nonzero Steklov eigenvalue on BRB_{R}. Suppose Robin parameter α[σ1(BR),0)\alpha\in[-\sigma_{1}(B_{R}),0). Then

  1. (1)

    for the function g(r)g(r), characterized in Proposition 3.1, we have

    g(r)>0, and g(r)g(r)α\displaystyle g^{\prime}(r)>0,\text{\quad and\quad}\frac{g^{\prime}(r)}{g(r)}\geq-\alpha (3.12)

    for r(0,R]r\in(0,R];

  2. (2)

    the second Robin eigenvalue

    λ2,α(BR)0.\lambda_{2,\alpha}(B_{R})\geq 0.
Proof.

Recall from (2.1) and (2.2) that SectM{1,4}\operatorname{Sect}_{M}\in\{-1,-4\} and

λ1(SR)=m1sinh2Rk1cosh2R<m1R2=λ1(S¯R),\lambda_{1}(S_{R})=\frac{m-1}{\sinh^{2}R}-\frac{k-1}{\cosh^{2}R}<\frac{m-1}{R^{2}}=\lambda_{1}(\bar{S}_{R}),

where S¯R\bar{S}_{R} is the boundary of the round ball B¯R\bar{B}_{R} of radius RR in m\mathbb{R}^{m}, then Escobar’s comparison [13, Theorem 2] gives

σ1(BR)σ1(B¯R)=1R.\displaystyle\sigma_{1}(B_{R})\leq\sigma_{1}(\bar{B}_{R})=\frac{1}{R}. (3.13)

Proof of (1). Let N(r)=J(r)g(r)N(r)=J(r)g^{\prime}(r). Using equation (3.8), we deduce that

N(r)=(λ1(Sr)λ2,α(BR))J(r)g(r).\displaystyle N^{\prime}(r)=\left(\lambda_{1}(S_{r})-\lambda_{2,\alpha}(B_{R})\right)J(r)g(r).

Recall

λ1(Sr)=mksinh2r+k1sinh2rcosh2r\lambda_{1}(S_{r})=\frac{m-k}{\sinh^{2}r}+\frac{k-1}{\sinh^{2}r\cosh^{2}r}

increasing in rr, then we have that N(r)N^{\prime}(r) has at most one zero in (0,R](0,R] and is positive near 0. Since N(0)=0N(0)=0 and N(R)=αJ(R)g(R)>0N(R)=-\alpha J(R)g^{\prime}(R)>0, then N(r)>0N(r)>0 for r(0,R]r\in(0,R], hence g(r)>0g^{\prime}(r)>0.

Now we turn to prove

g(r)g(r)α,r(0,R].\displaystyle\frac{g^{\prime}(r)}{g(r)}\geq-\alpha,\quad r\in(0,R]. (3.14)

Set v(r)=g(r)g(r)v(r)=\frac{g^{\prime}(r)}{g(r)}, then v(R)=αv(R)=-\alpha, v(r)>0v(r)>0 for r(0,R]r\in(0,R] and limr0+v(r)=+\lim_{r\rightarrow 0^{+}}v(r)=+\infty. Using H(r)=λ1(Sr)H^{\prime}(r)=-\lambda_{1}(S_{r}) and rewriting equation (3.8) as an ODE for vv yields

v+v2+Hv+(λ2,α(BR)+H)=0.\displaystyle v^{\prime}+v^{2}+Hv+\left(\lambda_{2,\alpha}(B_{R})+H^{\prime}\right)=0. (3.15)

Suppose (3.14) is not true, then there exists r0(0,R)r_{0}\in(0,R) such that

v(r0)=0, v′′(r0)0, and v(r0)<α.\displaystyle v^{\prime}(r_{0})=0,\text{\quad\quad}v^{\prime\prime}(r_{0})\geq 0,\text{\quad and\quad}v(r_{0})<-\alpha.

On one hand, by differentiating (3.15) in rr, we have that at r=r0r=r_{0}

0\displaystyle 0 =v′′(r0)+H(r0)v(r0)+H′′(r0)\displaystyle=v^{\prime\prime}(r_{0})+H^{\prime}(r_{0})v(r_{0})+H^{\prime\prime}(r_{0})
>αH(r0)+H′′(r0).\displaystyle>-\alpha H^{\prime}(r_{0})+H^{\prime\prime}(r_{0}). (3.16)

On the other hand, via direct calculations, we have

H′′H=\displaystyle\frac{H^{\prime\prime}}{H^{\prime}}= 2cothr2(k1)tanhr(mk)cosh2r+k1\displaystyle-2\coth r-2(k-1)\frac{\tanh r}{(m-k)\cosh^{2}r+k-1}
<\displaystyle< 2cothR\displaystyle-2\coth R
<\displaystyle< α,\displaystyle\alpha,

where in the last inequality we used assumption α[σ1(BR),0)\alpha\in[-\sigma_{1}(B_{R}),0) and inequality (3.13). In particular, we have

αH(r0)+H′′(r0)0,-\alpha H^{\prime}(r_{0})+H^{\prime\prime}(r_{0})\geq 0,

contradicting with (3). Hence inequality (3.14) comes true.

Proof of (2). Since

BRg(r)ψi(θ)𝑑A=0,1im,\displaystyle\int_{\partial B_{R}}g(r)\psi_{i}(\theta)\,dA=0,\quad 1\leq i\leq m,

the functions ui=g(r)ψi(θ)u_{i}=g(r)\psi_{i}(\theta) (1im1\leq i\leq m) are test functions for σ1(BR)\sigma_{1}(B_{R}). Where ψi\psi_{i}’s are the linear coordinate functions restricted to 𝕊m1\mathbb{S}^{m-1}. Therefore, by (3.3) we get

i=1mBR|ui|2𝑑μ\displaystyle\sum\limits_{i=1}^{m}\int_{B_{R}}|\nabla u_{i}|^{2}\,d\mu σ1(BR)i=1mBR|ui|2𝑑A\displaystyle\geq\sigma_{1}(B_{R})\sum\limits_{i=1}^{m}\int_{\partial B_{R}}|u_{i}|^{2}\,dA
αi=1mBR|ui|2𝑑A.\displaystyle\geq-\alpha\sum\limits_{i=1}^{m}\int_{\partial B_{R}}|u_{i}|^{2}\,dA.

Recall that

λ2,α(BR)\displaystyle\lambda_{2,\alpha}(B_{R}) =i=1mBR|ui|2𝑑μ+αi=1mBR|ui|2𝑑Ai=1mBRui2𝑑μ,\displaystyle=\frac{\sum\limits_{i=1}^{m}\int_{B_{R}}|\nabla u_{i}|^{2}\,d\mu+\alpha\sum\limits_{i=1}^{m}\int_{\partial B_{R}}|u_{i}|^{2}\,dA}{\sum\limits_{i=1}^{m}\int_{B_{R}}u_{i}^{2}\,d\mu},

so λ2,α(BR)0.\lambda_{2,\alpha}(B_{R})\geq 0. \square

4. A monotonicity lemma

Relabelling the solution to (3.8) as g1g_{1}, we define (with a slight abuse of notation) the function g:[0,)[0,)g:[0,\infty)\to[0,\infty) by

g(r):={g1(r),rR,g1(R)eα(rR),r>R.g(r):=\begin{cases}g_{1}(r),&r\leq R,\\ g_{1}(R)e^{-\alpha(r-R)},&r>R.\end{cases} (4.1)

By definition, gg is continuously differentiable and non-decreasing on (0,)(0,\infty). The following monotonicity lemma plays a key role in the proof of Theorem 1.2.

Lemma 4.1.

Assume that α[σ1(BR),0]\alpha\in[-\sigma_{1}(B_{R}),0]. Define F:[0,)F:[0,\infty)\to\mathbb{R} by

F(r):=g(r)2H(r)g2(r)+2αg(r)g(r)+αH(r)g2(r),F(r):=g^{\prime}(r)^{2}-H^{\prime}(r)g^{2}(r)+2\alpha g(r)g^{\prime}(r)+\alpha H(r)g^{2}(r), (4.2)

where g(r)g(r) is defined in (4.1). Then FF is monotonically decreasing on (0,)(0,\infty), and

F(r)78(mksinh3r+k1sinh3rcosh3r)g2(R)F^{\prime}(r)\leq-\frac{7}{8}\big{(}\frac{m-k}{\sinh^{3}r}+\frac{k-1}{\sinh^{3}r\cosh^{3}r}\big{)}g^{2}(R) (4.3)

for r>Rr>R.

Proof.

Case 1. 0<rR0<r\leq R. Differentiating FF in rr yields

F(r)\displaystyle F^{\prime}(r) =2g(r)g′′(r)H′′(r)g2(r)2H(r)g(r)g(r)I\displaystyle=\underbrace{2g^{\prime}(r)g^{\prime\prime}(r)-H^{\prime\prime}(r)g^{2}(r)-2H^{\prime}(r)g(r)g^{\prime}(r)}_{I}
+2αg(r)2+2αg(r)g′′(r)+αH(r)g2(r)+2αH(r)g(r)g(r)II.\displaystyle\quad+\underbrace{2\alpha g^{\prime}(r)^{2}+2\alpha g(r)g^{\prime\prime}(r)+\alpha H^{\prime}(r)g^{2}(r)+2\alpha H(r)g(r)g^{\prime}(r)}_{II}.

Using the ODE (3.8) of g(r)g(r), we have

I\displaystyle I =2g(r)g′′(r)H′′(r)g2(r)2H(r)g(r)g(r)\displaystyle=2g^{\prime}(r)g^{\prime\prime}(r)-H^{\prime\prime}(r)g^{2}(r)-2H^{\prime}(r)g(r)g^{\prime}(r)
=2H(r)g(r)2H′′(r)g2(r)4H(r)g(r)g(r)2λ2,α(BR)g(r)g(r)\displaystyle=-2H(r)g^{\prime}(r)^{2}-H^{\prime\prime}(r)g^{2}(r)-4H^{\prime}(r)g(r)g^{\prime}(r)-2\lambda_{2,\alpha}(B_{R})g(r)g^{\prime}(r)
=2((m1)cothr+(k1)tanhr)g(r)2+4(mksinh2r+k1sinh2rcosh2r)g(r)g(r)\displaystyle=-2\big{(}(m-1)\coth r+(k-1)\tanh r\big{)}g^{\prime}(r)^{2}+4\big{(}\frac{m-k}{\sinh^{2}r}+\frac{k-1}{\sinh^{2}r\cosh^{2}r}\big{)}g(r)g^{\prime}(r)
2((mk)coshrsinh3r+(k1)cosh2r+sinh2rsinh3rcosh3r)g(r)22λ2,α(BR)g(r)g(r).\displaystyle\quad-2\big{(}(m-k)\frac{\cosh r}{\sinh^{3}r}+(k-1)\frac{\cosh^{2}r+\sinh^{2}r}{\sinh^{3}r\cosh^{3}r}\big{)}g(r)^{2}-2\lambda_{2,\alpha}(B_{R})g(r)g^{\prime}(r).

Since

2(mksinh2r+k1sinh2rcosh2r)gg((m1)cothr+(k1)tanhr)g(r)2\displaystyle\quad 2\big{(}\frac{m-k}{\sinh^{2}r}+\frac{k-1}{\sinh^{2}r\cosh^{2}r}\big{)}gg^{\prime}-\big{(}(m-1)\coth r+(k-1)\tanh r\big{)}g^{\prime}(r)^{2}
((mk)coshrsinh3r+(k1)cosh2r+sinh2rsinh3rcosh3r)g2\displaystyle\quad-\big{(}(m-k)\frac{\cosh r}{\sinh^{3}r}+(k-1)\frac{\cosh^{2}r+\sinh^{2}r}{\sinh^{3}r\cosh^{3}r}\big{)}g^{2}
=(mk)(coshrsinhrg(r)22sinh2rgg+coshrsinh3rg2)\displaystyle=-(m-k)\big{(}\frac{\cosh r}{\sinh r}g^{\prime}(r)^{2}-\frac{2}{\sinh^{2}r}gg^{\prime}+\frac{\cosh r}{\sinh^{3}r}g^{2}\big{)}
(k1)(cothr+tanhr)g(r)2+2(k1)(1sinh2rcosh2r)gg\displaystyle\quad-(k-1)(\coth r+\tanh r)g^{\prime}(r)^{2}+2(k-1)\big{(}\frac{1}{\sinh^{2}r\cosh^{2}r}\big{)}gg^{\prime}
(k1)(cosh2r+sinh2rsinh3rcosh3r)g2\displaystyle\quad-(k-1)\big{(}\frac{\cosh^{2}r+\sinh^{2}r}{\sinh^{3}r\cosh^{3}r}\big{)}g^{2}
(k1)(sinh2r+cosh2rsinhrcoshrg(r)22ggsinh2rcosh2r+sinh2r+cosh2rsinh3rcosh3rg2)\displaystyle\leq-(k-1)\big{(}\frac{\sinh^{2}r+\cosh^{2}r}{\sinh r\cosh r}g^{\prime}(r)^{2}-\frac{2gg^{\prime}}{\sinh^{2}r\cosh^{2}r}+\frac{\sinh^{2}r+\cosh^{2}r}{\sinh^{3}r\cosh^{3}r}g^{2}\big{)}
<0,\displaystyle<0,

then we have

I2λ2,α(BR)g(r)g(r).\displaystyle I\leq-2\lambda_{2,\alpha}(B_{R})g(r)g^{\prime}(r). (4.4)

Using (3.8) again, we estimate

II\displaystyle II =2α(g)2+2αgg′′+αHg2+2αHgg\displaystyle=2\alpha(g^{\prime})^{2}+2\alpha gg^{\prime\prime}+\alpha H^{\prime}g^{2}+2\alpha Hgg^{\prime}
=2α(g)2αHg22αλ2,α(BR)g2\displaystyle=2\alpha(g^{\prime})^{2}-\alpha H^{\prime}g^{2}-2\alpha\lambda_{2,\alpha}(B_{R})g^{2}
=α(2(g)2+(mksinh2r+k1sinh2rcosh2r)g22λ2,α(BR)g2)\displaystyle=\alpha\left(2(g^{\prime})^{2}+(\frac{m-k}{\sinh^{2}r}+\frac{k-1}{\sinh^{2}r\cosh^{2}r})g^{2}-2\lambda_{2,\alpha}(B_{R})g^{2}\right)
2αλ2,α(BR)g2.\displaystyle\leq-2\alpha\lambda_{2,\alpha}(B_{R})g^{2}. (4.5)

Combing (4.4) and (4) together, we get

F(r)\displaystyle F^{\prime}(r) =I+II\displaystyle=I+II
<2λ2,α(BR)gg2αλ2,α(BR)g2\displaystyle<-2\lambda_{2,\alpha}(B_{R})gg^{\prime}-2\alpha\lambda_{2,\alpha}(B_{R})g^{2}
=2λ2,α(BR)g(r)(g(r)+αg(r))\displaystyle=-2\lambda_{2,\alpha}(B_{R})g(r)\left(g^{\prime}(r)+\alpha g(r)\right)
0,\displaystyle\leq 0,

where in the last inequality we used gαgg^{\prime}\geq-\alpha g on (0,R](0,R] and λ2,α(BR)0\lambda_{2,\alpha}(B_{R})\geq 0. Therefore, F(r)F(r) is monotonically decreasing on (0,R](0,R].

Case 2. rRr\geq R. Recall from (4.1) that g(r)=g(R)eα(rR)g(r)=g(R)e^{-\alpha(r-R)}, so we have

F(r)=(α2H(r)+αH(r))g2(r).\displaystyle F(r)=\left(-\alpha^{2}-H^{\prime}(r)+\alpha H(r)\right)g^{2}(r).

Differentiating FF in rr yields

F(r)=(2α3H′′(r)+3αH(r)2α2H(r))g2(r).F^{\prime}(r)=\left(2\alpha^{3}-H^{\prime\prime}(r)+3\alpha H^{\prime}(r)-2\alpha^{2}H(r)\right)g^{2}(r).

Since

H′′(r)3αH(r)+2α2H(r)\displaystyle\quad H^{\prime\prime}(r)-3\alpha H^{\prime}(r)+2\alpha^{2}H(r)
=2mksinh3r+2(k1)sinh2r+cosh2rcosh3rsinh3r+3α(mksinh2r+k1sinh2rcosh2r)\displaystyle=2\frac{m-k}{\sinh^{3}r}+2(k-1)\frac{\sinh^{2}r+\cosh^{2}r}{\cosh^{3}r\sinh^{3}r}+3\alpha(\frac{m-k}{\sinh^{2}r}+\frac{k-1}{\sinh^{2}r\cosh^{2}r})
+2α2((mk)coshrsinhr+(k1)sinh2r+cosh2rsinhrcoshr)\displaystyle\quad+2\alpha^{2}\left((m-k)\frac{\cosh r}{\sinh r}+(k-1)\frac{\sinh^{2}r+\cosh^{2}r}{\sinh r\cosh r}\right)
mksinhr(2sinh2r+3αsinhr+2α2)\displaystyle\geq\frac{m-k}{\sinh r}\left(\frac{2}{\sinh^{2}r}+\frac{3\alpha}{\sinh r}+2\alpha^{2}\right)
+k1coshrsinhr(2sinh2rcosh2r+3αsinhrcoshr+2α2)\displaystyle\quad+\frac{k-1}{\cosh r\sinh r}\left(\frac{2}{\sinh^{2}r\cosh^{2}r}+\frac{3\alpha}{\sinh r\cosh r}+2\alpha^{2}\right)
78(mksinh3r+k1sinh3rcosh3r),\displaystyle\geq\frac{7}{8}\big{(}\frac{m-k}{\sinh^{3}r}+\frac{k-1}{\sinh^{3}r\cosh^{3}r}\big{)},

then

F(r)\displaystyle F^{\prime}(r) 78(mksinh3r+k1sinh3rcosh3r)g2(r)\displaystyle\leq-\frac{7}{8}\big{(}\frac{m-k}{\sinh^{3}r}+\frac{k-1}{\sinh^{3}r\cosh^{3}r}\big{)}g^{2}(r)
78(mksinh3r+k1sinh3rcosh3r)g2(R),\displaystyle\leq-\frac{7}{8}\big{(}\frac{m-k}{\sinh^{3}r}+\frac{k-1}{\sinh^{3}r\cosh^{3}r}\big{)}g^{2}(R),

proving (4.3), and F(r)F(r) is monotonically decreasing on [R,+)[R,+\infty).

\square

5. Proof of Theorem 1.2

In this section, we shall prove Theorem 1.2. Before doing this, we first recall the following centre of mass lemma, which will be used later.

Lemma 5.1.

There exists a point pp in the closed geodesic convex hull of Ω\Omega, such that

Ωg(rp(x))expp1(x)rp(x)u1(x)𝑑μ=0,\int_{\Omega}g(r_{p}(x))\frac{\exp_{p}^{-1}(x)}{r_{p}(x)}u_{1}(x)\,d\mu=0, (5.1)

where gg is defined in (4.1), rp(x)=distg(p,x)r_{p}(x)=\operatorname{dist}_{g}(p,x), expp1\exp_{p}^{-1} is the inverse of the exponential map expp:TpMM\exp_{p}:T_{p}M\to M, and u1u_{1} is a first positive eigenfunction for λ1,α(Ω)\lambda_{1,\alpha}(\Omega).

Proof.

This lemma can be proved by Brouwer’s fixed point theorem (cf. [21, Page 635]). Here we give a more direct proof by following arguments in [16, Prop. 3] and [17, Sect. 7.4.3]. Let

G(y)=Ω(0ry(x)g(t)𝑑t)u1(x)𝑑μx,yM.G(y)=\int_{\Omega}\big{(}\int_{0}^{r_{y}(x)}g(t)\,dt\big{)}u_{1}(x)\,d\mu_{x},\quad\quad y\in M.

Since gg is monotone increasing on (0,)(0,\infty) and u1(x)u_{1}(x) is positive on Ω\Omega, then G(y)G(y) attains its minimum at pp in the convex hull of Ω\Omega, hence

0=G(p)=Ωg(rp(x))expp1(x)rp(x)u1(x)𝑑μ,0=\nabla G(p)=\int_{\Omega}g(r_{p}(x))\frac{\exp_{p}^{-1}(x)}{r_{p}(x)}u_{1}(x)\,d\mu,

proving the lemma. \square

Now we turn to prove Theorem 1.2.

Proof of Theorem 1.2.

From here on, we fix the point pp according to Lemma 5.1 so that (5.1) holds. Let (r,θ)(r,\theta) denote the polar coordinates centered at pp and SrS_{r} denote the distance sphere with respect to pp. Then we have

expp1(x)rp(x)=(ψ1(θ),ψ2(θ),,ψm(θ)),\displaystyle\frac{\exp^{-1}_{p}(x)}{r_{p}(x)}=\left(\psi_{1}(\theta),\psi_{2}(\theta),\cdots,\psi_{m}(\theta)\right),

where ψi\psi_{i}’s are the restrictions of the linear coordinate functions on 𝕊m1\mathbb{S}^{m-1}.

For 1im1\leq i\leq m, we define

vi(x):=g(rp(x))ψi(θ),\displaystyle v_{i}(x):=g(r_{p}(x))\psi_{i}(\theta),

and rewrite (5.1) as

Ωvi(x)u1(x)𝑑μ=0.\displaystyle\int_{\Omega}v_{i}(x)u_{1}(x)\,d\mu=0.

So viv_{i}’s are test functions for λ2,α(Ω)\lambda_{2,\alpha}(\Omega) and we conclude

λ2,α(Ω)i=1mΩ|vi|2𝑑μ+αi=1mΩvi2𝑑Ai=1mΩvi2𝑑μ.\displaystyle\lambda_{2,\alpha}(\Omega)\leq\frac{\sum\limits_{i=1}^{m}\int_{\Omega}|\nabla v_{i}|^{2}\,d\mu+\alpha\sum\limits_{i=1}^{m}\int_{\partial\Omega}v_{i}^{2}\,dA}{\sum\limits_{i=1}^{m}\int_{\Omega}v_{i}^{2}\,d\mu}. (5.2)

Using (2.3) and i=1mψi2=1\sum_{i=1}^{m}\psi_{i}^{2}=1, we compute that

i=1mΩ|vi|2𝑑μ\displaystyle\sum_{i=1}^{m}\int_{\Omega}\left|\nabla v_{i}\right|^{2}\,d\mu =i=1mΩ|(g(rp)ψi)|2𝑑μ\displaystyle=\sum_{i=1}^{m}\int_{\Omega}\left|\nabla\left(g(r_{p})\psi_{i}\right)\right|^{2}\,d\mu
=i=1mΩ|g(rp)|2ψi2+g2(rp)|Srψi|2dμ\displaystyle=\sum_{i=1}^{m}\int_{\Omega}\left|g^{\prime}(r_{p})\right|^{2}\psi^{2}_{i}+g^{2}(r_{p})|\nabla^{S_{r}}\psi_{i}|^{2}\,d\mu
=Ω|g(rp)|2H(rp)g2(rp)dμ,\displaystyle=\int_{\Omega}\left|g^{\prime}(r_{p})\right|^{2}-H^{\prime}(r_{p})g^{2}(r_{p})\,d\mu, (5.3)

and

i=1mΩvi2𝑑μ=i=1mΩ|g(rp)|2ψi2𝑑μ=Ω|g(rp)|2𝑑μ,\sum_{i=1}^{m}\int_{\Omega}v_{i}^{2}\,d\mu=\sum_{i=1}^{m}\int_{\Omega}|g(r_{p})|^{2}\psi_{i}^{2}\,d\mu=\int_{\Omega}|g(r_{p})|^{2}\,d\mu, (5.4)

and

i=1mΩvi2𝑑A=i=1mΩ|g(rp)|2ψi2𝑑A=Ω|g(rp)|2𝑑A.\sum_{i=1}^{m}\int_{\partial\Omega}v_{i}^{2}\,dA=\sum_{i=1}^{m}\int_{\partial\Omega}|g(r_{p})|^{2}\psi_{i}^{2}\,dA=\int_{\partial\Omega}|g(r_{p})|^{2}\,dA. (5.5)

Using |rp|=1|\nabla r_{p}|=1 and Δrp=H(rp)\Delta r_{p}=H(r_{p}), we estimate

Ωg2(rp)𝑑A\displaystyle\int_{\partial\Omega}g^{2}(r_{p})\,dA Ωg2(rp)rp,ν𝑑A\displaystyle\geq\int_{\partial\Omega}g^{2}(r_{p})\langle\nabla r_{p},\nu\rangle\,dA
=Ωdiv(g2(rp)rp)𝑑μ\displaystyle=\int_{\Omega}\operatorname{div}\left(g^{2}(r_{p})\nabla r_{p}\right)\,d\mu
=Ω(g2)(rp)+g2(rp)Δrpdμ\displaystyle=\int_{\Omega}(g^{2})^{\prime}(r_{p})+g^{2}(r_{p})\Delta r_{p}\,d\mu
=Ω(g2)(rp)+H(rp)g2(rp)dμ,\displaystyle=\int_{\Omega}(g^{2})^{\prime}(r_{p})+H(r_{p})g^{2}(r_{p})\,d\mu, (5.6)

where ν\nu is the outward unit normal to Ω\partial\Omega. So substituting (5),(5.4) and (5.5) into (5.2) yields

λ2,α(Ω)Ω(|g(rp)|2H(rp)g2(rp)+2αg(rp)g(rp)+αH(rp)g2(rp))𝑑μΩg2(rp)𝑑μ,\displaystyle\lambda_{2,\alpha}(\Omega)\leq\frac{\int_{\Omega}\left(\left|g^{\prime}(r_{p})\right|^{2}-H^{\prime}(r_{p})g^{2}(r_{p})+2\alpha g(r_{p})g^{\prime}(r_{p})+\alpha H(r_{p})g^{2}(r_{p})\right)\,d\mu}{\int_{\Omega}g^{2}(r_{p})\,d\mu},

where we used (5) and the fact α0\alpha\leq 0. That is

λ2,α(Ω)ΩF(rp)𝑑μΩg2(rp)𝑑μ.\lambda_{2,\alpha}(\Omega)\leq\frac{\int_{\Omega}F(r_{p})\,d\mu}{\int_{\Omega}g^{2}(r_{p})\,d\mu}. (5.7)

Let BRB_{R} be the geodesic ball of radius RR, having the same volume as Ω\Omega’s and centered at pp so that (5.1) holds. Recall that g(r)ψi(θ)g(r)\psi_{i}(\theta) are the eigenfunctions corresponding to λ2,α(BR)\lambda_{2,\alpha}(B_{R}), so then

λ2,α(BR)=BR|g|2(rp)H(rp)g2(rp)dμ+αBRg2(rp)𝑑ABRg2(rp)𝑑μ\displaystyle\lambda_{2,\alpha}(B_{R})=\frac{\int_{B_{R}}|g^{\prime}|^{2}(r_{p})-H^{\prime}(r_{p})g^{2}(r_{p})\,d\mu+\alpha\int_{\partial B_{R}}g^{2}(r_{p})\,dA}{\int_{B_{R}}g^{2}(r_{p})\,d\mu}

and

BRg2(rp)𝑑A\displaystyle\int_{\partial B_{R}}g^{2}(r_{p})\,dA =BRg2(rp)rp,ν𝑑A\displaystyle=\int_{\partial B_{R}}\langle g^{2}(r_{p})\nabla r_{p},\nu\rangle\,dA
=BRdiv(g2(rp)rp)𝑑μ\displaystyle=\int_{B_{R}}\operatorname{div}\left(g^{2}(r_{p})\nabla r_{p}\right)\,d\mu
=BR(g2)+g2Δrpdμ\displaystyle=\int_{B_{R}}(g^{2})^{\prime}+g^{2}\Delta r_{p}\,d\mu
=BR(g2)+H(rp)g2dμ.\displaystyle=\int_{B_{R}}(g^{2})^{\prime}+H(r_{p})g^{2}\,d\mu.

Therefore we conclude

λ2,α(BR)=BRF(rp)𝑑μBRg2(rp)𝑑μ.\displaystyle\lambda_{2,\alpha}(B_{R})=\frac{\int_{B_{R}}F(r_{p})\,d\mu}{\int_{B_{R}}g^{2}(r_{p})\,d\mu}. (5.8)

Recall that gg defined in (4.1) is increasing, we have

Ωg2(rp)𝑑μ\displaystyle\int_{\Omega}g^{2}(r_{p})\,d\mu =ΩBRg2(rp)𝑑μ+ΩBRg2(rp)𝑑μ\displaystyle=\int_{\Omega\cap B_{R}}g^{2}(r_{p})\,d\mu+\int_{\Omega\setminus B_{R}}g^{2}(r_{p})\,d\mu
=BRg2(rp)𝑑μ+ΩBRg2(rp)𝑑μBRΩg2(rp)𝑑μ\displaystyle=\int_{B_{R}}g^{2}(r_{p})\,d\mu+\int_{\Omega\setminus B_{R}}g^{2}(r_{p})\,d\mu-\int_{B_{R}\setminus\Omega}g^{2}(r_{p})\,d\mu
BRg2(rp)𝑑μ.\displaystyle\geq\int_{B_{R}}g^{2}(r_{p})\,d\mu. (5.9)

On the other hand, by Lemma 4.1, FF is monotonically decreasing, so then

ΩF(rp)𝑑μ\displaystyle\int_{\Omega}F(r_{p})\,d\mu =BRF(rp)𝑑μ+ΩBRF(rp)𝑑μBRΩF(rp)𝑑μ\displaystyle=\int_{B_{R}}F(r_{p})\,d\mu+\int_{\Omega\setminus B_{R}}F(r_{p})\,d\mu-\int_{B_{R}\setminus\Omega}F(r_{p})\,d\mu
BRF(rp)𝑑μ+ΩBRF(rp)F(R)dμ\displaystyle\leq\int_{B_{R}}F(r_{p})\,d\mu+\int_{\Omega\setminus B_{R}}F(r_{p})-F(R)\,d\mu
BRF(rp)𝑑μ+BR1BRF(rp)F(R)dμ,\displaystyle\leq\int_{B_{R}}F(r_{p})\,d\mu+\int_{B_{R_{1}}\setminus B_{R}}F(r_{p})-F(R)\,d\mu, (5.10)

where BR1B_{R_{1}} is the ball of radius R1R_{1} centered at pp satisfying

|BR1BR|=|ΩBR|.|B_{R_{1}}\setminus B_{R}|=|\Omega\setminus B_{R}|.

Using (4.3), we estimate that

BR1BRF(rp)F(R)dμ\displaystyle\quad\int_{B_{R_{1}}\setminus B_{R}}F(r_{p})-F(R)\,d\mu
BR1BR78(mksinh3R1+k1sinh3R1cosh3R1)g2(R)(rpR)𝑑μ\displaystyle\leq-\int_{B_{R_{1}}\setminus B_{R}}\frac{7}{8}\big{(}\frac{m-k}{\sinh^{3}R_{1}}+\frac{k-1}{\sinh^{3}R_{1}\cosh^{3}R_{1}}\big{)}g^{2}(R)(r_{p}-R)\,d\mu
C(R1R)2,\displaystyle\leq-C\big{(}R_{1}-R\big{)}^{2}, (5.11)

here and thereafter, CC denotes a constant depending on mm, kk and |Ω|=v|\Omega|=v, which may change from line to line. Inequalities (5) and (5) yields

ΩF(rp)𝑑μBRF(rp)𝑑μC|ΩBR|2,\int_{\Omega}F(r_{p})\,d\mu\leq\int_{B_{R}}F(r_{p})\,d\mu-C|\Omega\setminus B_{R}|^{2}, (5.12)

where we used |BR1BR|=O((R1R)2)|B_{R_{1}}\setminus B_{R}|=O((R_{1}-R)^{2}).

From (5.7), (5.8), (5) and (5.12), we deduce

λ2,α(Ω)\displaystyle\lambda_{2,\alpha}(\Omega) BRF(rp)𝑑μC|ΩBR|2BRg2(rp)𝑑μ\displaystyle\leq\frac{\int_{B_{R}}F(r_{p})\,d\mu-C|\Omega\setminus B_{R}|^{2}}{\int_{B_{R}}g^{2}(r_{p})\,d\mu}
=λ2,α(BR)C|ΩBR|2,\displaystyle=\lambda_{2,\alpha}(B_{R})-C|\Omega\setminus B_{R}|^{2},

proving inequality (1.3).

Equality in (1.3) clearly holds if Ω=B\Omega=B. Since we always have λ2,α(Ω)λ2,α(B)\lambda_{2,\alpha}(\Omega)\leq\lambda_{2,\alpha}(B), equality in (1.3) implies 𝒜(Ω)=0\mathcal{A}(\Omega)=0, i.e. Ω=B\Omega=B.

Therefore, Theorem 1.2 is proved. \square

References

  • [1] A. R. Aithal and G. Santhanam. Sharp upper bound for the first non-zero Neumann eigenvalue for bounded domains in rank-11 symmetric spaces. Trans. Amer. Math. Soc., 348(10):3955–3965, 1996.
  • [2] Mark S. Ashbaugh and Rafael D. Benguria. Sharp upper bound to the first nonzero Neumann eigenvalue for bounded domains in spaces of constant curvature. J. London Math. Soc. (2), 52(2):402–416, 1995.
  • [3] Miriam Bareket. On an isoperimetric inequality for the first eigenvalue of a boundary value problem. SIAM J. Math. Anal., 8(2):280–287, 1977.
  • [4] Binoy and G. Santhanam. Sharp upperbound and a comparison theorem for the first nonzero Steklov eigenvalue. J. Ramanujan Math. Soc., 29(2):133–154, 2014.
  • [5] Marie-Hélène Bossel. Membranes élastiquement liées: extension du théorème de Rayleigh-Faber-Krahn et de l’inégalité de Cheeger. C. R. Acad. Sci. Paris Sér. I Math., 302(1):47–50, 1986.
  • [6] Lorenzo Brasco, Guido De Philippis, and Berardo Ruffini. Spectral optimization for the Stekloff-Laplacian: the stability issue. J. Funct. Anal., 262(11):4675–4710, 2012.
  • [7] Lorenzo Brasco and Aldo Pratelli. Sharp stability of some spectral inequalities. Geom. Funct. Anal., 22(1):107–135, 2012.
  • [8] F. Brock. An isoperimetric inequality for eigenvalues of the Stekloff problem. ZAMM Z. Angew. Math. Mech., 81(1):69–71, 2001.
  • [9] Dorin Bucur and Alessandro Giacomini. A variational approach to the isoperimetric inequality for the Robin eigenvalue problem. Arch. Ration. Mech. Anal., 198(3):927–961, 2010.
  • [10] Dorin Bucur and Alessandro Giacomini. Faber-Krahn inequalities for the Robin-Laplacian: a free discontinuity approach. Arch. Ration. Mech. Anal., 218(2):757–824, 2015.
  • [11] Philippe Castillon and Berardo Ruffini. A spectral characterization of geodesic balls in non-compact rank one symmetric spaces. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 19(4):1359–1388, 2019.
  • [12] Daniel Daners. A Faber-Krahn inequality for Robin problems in any space dimension. Math. Ann., 335(4):767–785, 2006.
  • [13] José F. Escobar. A comparison theorem for the first non-zero Steklov eigenvalue. J. Funct. Anal., 178(1):143–155, 2000.
  • [14] Pedro Freitas and David Krejčiřik. The first Robin eigenvalue with negative boundary parameter. Adv. Math., 280:322–339, 2015.
  • [15] Pedro Freitas and Richard S. Laugesen. From Steklov to Neumann and beyond, via Robin: the Szegö way. Canad. J. Math., 72(4):1024–1043, 2020.
  • [16] Pedro Freitas and Richard S. Laugesen. From Neumann to Steklov and beyond, via Robin: the Weinberger way. Amer. J. Math., 143(3):969–994, 2021.
  • [17] Antoine Henrot, editor. Shape optimization and spectral theory. De Gruyter Open, Warsaw, 2017.
  • [18] James Kennedy. An isoperimetric inequality for the second eigenvalue of the Laplacian with Robin boundary conditions. Proc. Amer. Math. Soc., 137(2):627–633, 2009.
  • [19] Xiaolong Li, Kui Wang, and Haotian Wu. On the second Robin eigenvalue of the Laplacian. Preprint, 2020. arXiv:2003.03087 [math.DG].
  • [20] G. Szegö. Inequalities for certain eigenvalues of a membrane of given area. J. Rational Mech. Anal., 3:343–356, 1954.
  • [21] H. F. Weinberger. An isoperimetric inequality for the NN-dimensional free membrane problem. J. Rational Mech. Anal., 5:633–636, 1956.
  • [22] Robert Weinstock. Inequalities for a classical eigenvalue problem. J. Rational Mech. Anal., 3:745–753, 1954.