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

Steklov eigenvalue problem on subgraphs of integer lattices

Wen Han Wen Han: School of Mathematical Sciences, Fudan University, Shanghai 200433, China. wenhan122@foxmail.com  and  Bobo Hua Bobo Hua: School of Mathematical Sciences, LMNS, Fudan University, Shanghai 200433, China; Shanghai Center for Mathematical Sciences, Fudan University, Shanghai 200433, China bobohua@fudan.edu.cn
Abstract.

We study the eigenvalues of the Dirichlet-to-Neumann operator on a finite subgraph of the integer lattice n.{\mathbb{Z}}^{n}. We estimate the first n+1n+1 eigenvalues using the number of vertices of the subgraph. As a corollary, we prove that the first non-trivial eigenvalue of the Dirichlet-to-Neumann operator tends to zero as the number of vertices of the subgraph tends to infinity.

1. Introduction

Given a bounded smooth domain Ω\Omega in n,\mathbb{R}^{n}, we consider the Steklov (eigenvalue) problem on Ω.\Omega. For a smooth function φ\varphi on Ω,\partial\Omega, we denote by uφu_{\varphi} the harmonic extension of φ\varphi into Ω\Omega which satisfies

{Δuφ(x)=0,xΩ,uφ|Ω=φ.\left\{\begin{array}[]{ll}\Delta u_{\varphi}(x)=0,&x\in\Omega,\\ u_{\varphi}|_{\partial\Omega}=\varphi.&\end{array}\right.

The Steklov problem on Ω\Omega is the following eigenvalue problem,

Λφ:=uφn=λφonΩ,\Lambda\varphi:=\frac{\partial u_{\varphi}}{\partial n}=\lambda\varphi\quad\mathrm{on}\ \partial\Omega,

where nn denotes the unit outward normal of Ω.\partial\Omega. The operator Λ\Lambda is called the Dirichlet-to-Neumann (DtN in short) operator, which is also known as the voltage-current map in physics, see e.g. Calderón’s problem [Cal06]. As a pseudo-differential operator, the eigenvalues of DtN operator Λ,\Lambda, called Steklov eigenvalues, are discrete and can be ordered as

0=σ1(Ω)σ2(Ω)σk(Ω).0=\sigma_{1}(\Omega)\leq\sigma_{2}(\Omega)\leq\cdots\leq\sigma_{k}(\Omega)\leq\cdots\uparrow\infty.

The Steklov problem has been intensively studied for domains in Euclidean spaces and Riemannian manifolds in the literature, see e.g. [Wei54, Esc97, Esc99, Bro01, HPS08, WX09, SU90, CSG11, Has11, GP12, FL14, KKP14, GLS16, FS16, Mat17, FW17, YY17].

For n=2,n=2, Weinstock [Wei54] proved that for any planar simply-connected domain Ω\Omega with analytic boundary,

(1.1) σ2(Ω)2πL(Ω),\sigma_{2}(\Omega)\leq\frac{2\pi}{\mathrm{L}(\partial\Omega)},

where L(Ω)\mathrm{L}(\partial\Omega) denotes the length of the boundary Ω.\partial\Omega. The statement was generalized to domains with Lipschitz boundary by Girouard and Polterovich[GP10]. The higher dimensional generalization was proved by Bucur, Ferone, Nitsch and Trombetti [BFNT17]: for any bounded convex Ωn,\Omega\subset{\mathbb{R}}^{n}, n3,n\geq 3,

σ2(Ω)(area(Ω))1n1σ2(B)(area(B))1n1,\sigma_{2}(\Omega)(\mathrm{area}(\partial\Omega))^{\frac{1}{n-1}}\leq\sigma_{2}(B)(\mathrm{area}(\partial B))^{\frac{1}{n-1}},

where BB is the unit ball and area()\mathrm{area}(\cdot) denotes the area of (),(\cdot), and the equality holds only for balls, see also Fraser and Schoen [FS19] for related results.

The estimate (1.1) was improved by Hersch and Payne [HP68] to the following

(1.2) 1σ2(Ω)+1σ3(Ω)L(Ω)π.\frac{1}{\sigma_{2}(\Omega)}+\frac{1}{\sigma_{3}(\Omega)}\geq\frac{\mathrm{L}(\partial\Omega)}{\pi}.

Concerning with the Steklov eigenvalues and the volume of the domain, Brock proved the following result.

Theorem 1.1 (Theorem 3 in [Bro01]).

Let Ω\Omega be a smooth domain in n.{\mathbb{R}}^{n}. Then

i=2n+11σi(Ω)C(n)(vol(Ω))1n,\sum_{i=2}^{n+1}\frac{1}{\sigma_{i}(\Omega)}\geq C(n)(\mathrm{vol}(\Omega))^{\frac{1}{n}},

where vol(Ω)\mathrm{vol}(\Omega) denotes the volume of Ω,\Omega, and C(n)=nωn1nC(n)=n\omega_{n}^{-\frac{1}{n}} (here ωn\omega_{n} is the volume of the unit ball in n{\mathbb{R}}^{n}).

In this paper, we study the Steklov problem on graphs and prove a discrete analog for Brock’s result. The DtN operator on a subgraph of a graph was introduced by [HHW17, HM17] independently, see e.g. [HHW18, Per19] for more results. A graph G=(V,E)G=(V,E) consists of the set of vertices VV and the set of edges E.E. In this paper, we only consider simple, undirected graphs. Two vertices x,yx,y are called neighbours, denoted by xyx\sim y, if there is an edge ee connecting xx and y,y, i.e. e={x,y}E.e=\{x,y\}\in E. Integer lattice graphs are of particular interest which serve as the discrete counterparts of n.{\mathbb{R}}^{n}. We denote by n{\mathbb{Z}}^{n} the set of integer nn-tuples n.{\mathbb{R}}^{n}. The nn-dimensional integer lattice graph, still denoted by n,{\mathbb{Z}}^{n}, is the graph consisting of the set of vertices V=nV={\mathbb{Z}}^{n} and the set of edges

E={{x,y}:x,yn,i=1n|xiyi|=1}.E=\left\{\{x,y\}:x,y\in{\mathbb{Z}}^{n},\sum_{i=1}^{n}|x_{i}-y_{i}|=1\right\}.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. We denote by

δΩ:={xnΩ:yΩ,s.t.xy}\delta\Omega:=\{x\in{\mathbb{Z}}^{n}\setminus\Omega:\exists\ y\in\Omega,s.t.\ x\sim y\}

the vertex boundary of Ω.\Omega. Analogous to the continuous setting, one can define the DtN operator on Ω,\Omega,

Λ:δΩδΩ,\Lambda:{\mathbb{R}}^{\delta\Omega}\to{\mathbb{R}}^{\delta\Omega},

where δΩ{\mathbb{R}}^{\delta\Omega} denotes the set of functions on δΩ,\delta\Omega, see Section 2 for details. In this paper, we denote by |||\cdot| the cardinality of a set. Note that Λ\Lambda can be written as a symmetric matrix whose eigenvalues are ordered as

(1.3) 0=λ1(Ω)λ2(Ω)λN(Ω),whereN=|δΩ|.0=\lambda_{1}(\Omega)\leq\lambda_{2}(\Omega)\leq\cdots\leq\lambda_{N}(\Omega),\ \text{where}~N=|\delta\Omega|.

The following is our main result, which is a discrete analog of [Bro01, Theorem 3].

Theorem 1.2.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. Then

(1.4) i=2n+11λi(Ω)C1|Ω|1nC2|Ω|,\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)}\geq C_{1}|\Omega|^{\frac{1}{n}}-\frac{C_{2}}{|\Omega|},

where λi(Ω)\lambda_{i}(\Omega) are Steklov eigenvalues on Ω,\Omega, C1=(64n3ωn1n)1,C_{1}=(64n^{3}\omega_{n}^{\frac{1}{n}})^{-1}, and C2=132n.C_{2}=\frac{1}{32n}.

Remark 1.3.

By this theorem, for |Ω|(2C2C1)nn+1,|\Omega|\geq(\frac{2C_{2}}{C_{1}})^{\frac{n}{n+1}},

i=2n+11λi(Ω)12C1|Ω|1n,\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)}\geq\frac{1}{2}C_{1}|\Omega|^{\frac{1}{n}},

where (2C2C1)nn+1=(4n2ωn1n)nn+1(\frac{2C_{2}}{C_{1}})^{\frac{n}{n+1}}={(4n^{2}\omega_{n}^{\frac{1}{n}})^{\frac{n}{n+1}}} is of order O(n32),n.O(n^{\frac{3}{2}}),n\to\infty.

The main ingredient of the proof of Brock’s result is a weighted isoperimetric inequality, see Lemma 3.5 below, which depends on the rotational symmetry of the Euclidean spaces. As is well-known in the discrete theory, the symmetrization approaches do not work on n{\mathbb{Z}}^{n} due to the lack of rotational symmetry. In this paper, we follow Brock’s idea to bound the Steklov eigenvalues by the geometric quantities of the subset in the lattice n,{\mathbb{Z}}^{n}, and then estimate these discrete quantities by their counterparts in the Euclidean space n{\mathbb{R}}^{n}, for which we can apply the symmetrization approach. Through this process we obtain the quantitative estimate of discrete Steklov eigenvalues, but lose the sharpness of the constants C1C_{1} and C2C_{2} in our result.

As a corollary, we obtain the estimate for the first non-trivial eigenvalue of the DtN operator.

Corollary 1.4.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. Then

(1.5) λ2(Ω)nC1|Ω|1nC2|Ω|,\lambda_{2}(\Omega)\leq\frac{n}{C_{1}|\Omega|^{\frac{1}{n}}-\frac{C_{2}}{|\Omega|}},

where C1C_{1} and C2C_{2} are the constants in Theorem 1.2.

Remark 1.5.

Recently, the above corollary has been extended to the DtN operators on subgraphs of Cayley graphs for discrete groups of polynomial growth by Perrin [Per20].

This yields an interesting consequence that for any sequence of finite subsets in n,{\mathbb{Z}}^{n}, {Ωi}i=1,\{\Omega_{i}\}_{i=1}^{\infty}, satisfying |Ωi|,|\Omega_{i}|\to\infty,

(1.6) λ2(Ωi)0,i.\lambda_{2}(\Omega_{i})\to 0,\quad i\to\infty.

This is a rather special property for integer lattices, which does not hold for general infinite graphs. At the end of the paper, we give an example of an infinite graph of bounded degree, Example 3.7, in which there exist subsets Ωi\Omega_{i} with |Ωi||\Omega_{i}|\to\infty such that

λ2(Ωi)c>0,i.\lambda_{2}(\Omega_{i})\geq c>0,\quad\forall i.

Note that this graph locally behaves similarly as a homogeneous tree of degree 3,3, which is a discrete hyperbolic model.

The paper is organized as follows: In the next section, we recall some facts on DtN operators on graphs. In Section 3, we prove Theorem 1.2.

2. Preliminaries

Let G=(V,E)G=(V,E) be a simple, undirected graph. For two subsets A,BA,B of VV, we define

E(A,B):={{x,y}E:xA,yB,orviceversa},E(A,B):=\{\{x,y\}\in E:~x\in A,~y\in B,\ \mathrm{or\ vice\ versa}\},

which is a set of edges connecting a vertex in AA to another vertex in B.B. For a subset Ω\Omega of VV, we write Ωc:=VΩ.\Omega^{c}:=V\setminus\Omega. The edge boundary of Ω\Omega is defined as Ω:=E(Ω,Ωc),\partial\Omega:=E(\Omega,\Omega^{c}), and we set Ω¯:=ΩδΩ\bar{\Omega}:=\Omega\cup\delta\Omega.

For a finite subset SS of V,V, we denote by S{\mathbb{R}}^{S} the set of functions on S,S, by 2(S)\ell^{2}(S) the Hilbert space of functions on SS equipped with the inner product

φ,ψS=xSφ(x)ψ(x),φ,ψS.\langle\varphi,\psi\rangle_{S}=\sum_{x\in S}\varphi(x)\psi(x),\quad\varphi,\psi\in\mathbb{R}^{S}.

In the following, we will take S=Ω,S=\Omega, δΩ,\delta\Omega, or Ω¯,\bar{\Omega}, respectively. For uΩ¯,u\in{\mathbb{R}}^{\bar{\Omega}}, the Dirichlet energy of uu is defined as

DΩ(u):={x,y}E(Ω,Ω¯)(u(x)u(y))2.D_{\Omega}(u):=\sum_{\{x,y\}\in E(\Omega,\bar{\Omega})}(u(x)-u(y))^{2}.

The polarization of the Dirichlet energy is given by

(2.1) DΩ(u,v):={x,y}E(Ω,Ω¯)(u(x)u(y))(v(x)v(y)),u,vΩ¯.D_{\Omega}(u,v):=\sum_{\{x,y\}\in E(\Omega,\bar{\Omega})}(u(x)-u(y))(v(x)-v(y)),\ \forall u,v\in{\mathbb{R}}^{\bar{\Omega}}.

The Laplace operator on Ω\Omega is defined as

Δ:Ω¯\displaystyle\Delta:{\mathbb{R}}^{\bar{\Omega}} Ω\displaystyle\rightarrow{\mathbb{R}}^{\Omega}
u\displaystyle u Δu(x):=yV:yx(u(y)u(x)).\displaystyle\mapsto\Delta u(x):=\sum_{y\in V:y\sim x}(u(y)-u(x)).

The exterior normal derivative on vertex boundary set δΩ\delta\Omega is defined as

n:Ω¯\displaystyle\frac{\partial}{\partial n}:{\mathbb{R}}^{\bar{\Omega}} δΩ\displaystyle\rightarrow{\mathbb{R}}^{\delta\Omega}
u\displaystyle u un(x):=yΩ:yx(u(x)u(y)),xδΩ.\displaystyle\mapsto\frac{\partial u}{\partial n}(x):=\sum_{y\in\Omega:y\sim x}(u(x)-u(y)),\ x\in\delta\Omega.

The Dirichlet-to-Neumann (DtN in short) operator is defined as

Λ:δΩ\displaystyle\Lambda:{\mathbb{R}}^{\delta\Omega} δΩ\displaystyle\rightarrow{\mathbb{R}}^{\delta\Omega}
φ\displaystyle\varphi Λφ:=uφn,\displaystyle\mapsto\Lambda\varphi:=\frac{\partial u_{\varphi}}{\partial n},

where uφu_{\varphi} is the harmonic function on Ω\Omega whose Dirichlet boundary condition on δΩ\delta\Omega is given by φ.\varphi. We recall the well-known Green formula.

Theorem 2.1 (Lemma 2.1 in [HHW17]).

For φδΩ,\varphi\in\mathbb{R}^{\delta{\Omega}},

(2.2) DΩ(uφ)=Δuφ,uφΩ+uφn,uφδΩ,D_{\Omega}(u_{\varphi})=-\langle\Delta u_{\varphi},u_{\varphi}\rangle_{\Omega}+\langle\frac{\partial u_{\varphi}}{\partial n},u_{\varphi}\rangle_{\delta\Omega},

where uφu_{\varphi} is the harmonic extension of φ\varphi in Ω.\Omega.

Proposition 2.1.

The DtN operator Λ\Lambda can be represented as a symmetric matrix, and hence is diagonalizable. The multiplicity of zero-eigenvalues of Λ\Lambda is equal to the number of connected components of the graph (Ω¯,E(Ω,Ω¯))(\bar{\Omega},E(\Omega,\bar{\Omega})).

Proof.

Assume that ϕ,φδΩ\phi,\varphi\in\mathbb{R}^{\delta\Omega} are eigenfunctions, which can be regarded as vectors in N{\mathbb{R}}^{N} with N=|δΩ|,N=|\delta\Omega|, of the DtN operator Λ,\Lambda, which is a N×NN\times N matrix. Using the polarization of the Dirichlet energy (2.1) and Green formula (2.2), we have

Λuϕ,uφδΩ=DΩ(uϕ,uφ)=DΩ(uφ,uϕ)=uϕ,ΛuφδΩ.\langle\Lambda u_{\phi},u_{\varphi}\rangle_{\delta\Omega}=D_{\Omega}(u_{\phi},u_{\varphi})=D_{\Omega}(u_{\varphi},u_{\phi})=\langle u_{\phi},\Lambda u_{\varphi}\rangle_{\delta\Omega}.

Hence, Λ\Lambda is a symmetric matrix, and hence is diagonalizable.

Assume that φδΩ\varphi\in\mathbb{R}^{\delta\Omega} is an eigenfunction corresponding to the eigenvalue 0 of the DtN operator Λ,\Lambda, then by the Green formula (2.2),

DΩ(uφ)=Δuφ,uφΩ+uφn,uφδΩ=0.D_{\Omega}(u_{\varphi})=-\langle\Delta u_{\varphi},u_{\varphi}\rangle_{\Omega}+\langle\frac{\partial u_{\varphi}}{\partial n},u_{\varphi}\rangle_{\delta\Omega}=0.

Thus, uφu_{\varphi} is constant on every connected component of (Ω¯,E(Ω,Ω¯)).(\bar{\Omega},E(\Omega,\bar{\Omega})). This yields the result. ∎

Let λ\lambda be an eigenvalue of the DtN operator Λ\Lambda satisfying

uφn=λuφonδΩ.\frac{\partial u_{\varphi}}{\partial n}=\lambda u_{\varphi}\ \mathrm{on\ \delta\Omega}.

Then, by the Green formula (2.2) we get

DΩ(uφ)=λuφ,uφδΩ=λzδΩφ2(z).D_{\Omega}(u_{\varphi})=\lambda\langle u_{\varphi},u_{\varphi}\rangle_{\delta\Omega}=\lambda\sum_{z\in\delta\Omega}{\varphi}^{2}(z).

The following variational principle is useful, see e.g. [Ban80, p.99].

Theorem 2.2.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. Then

(2.3) i=2p1λi(Ω)=max{i=2pzδΩvi2(z):viΩ¯,DΩ(vi,vj)=δij,zδΩvi(z)=0, 2i,jp},\sum^{p}_{i=2}\frac{1}{\lambda_{i}(\Omega)}=\max\left\{\sum^{p}_{i=2}\sum_{z\in\delta\Omega}v_{i}^{2}(z):v_{i}\in{\mathbb{R}}^{\bar{\Omega}},D_{\Omega}(v_{i},v_{j})=\delta_{ij},\sum_{z\in\delta\Omega}v_{i}(z)=0,\ 2\leq i,j\leq p\right\},

where p2p\geq 2 and δij={1,i=j,0,ij,\delta_{ij}=\left\{\begin{array}[]{ll}1,&i=j,\\ 0,&i\neq j,\end{array}\right. where we adopt the traditional convention that λl=+ for l>|δΩ|\lambda_{l}=+\infty\text{ for }l>|\delta\Omega|.

3. The proof of Theorem 1.2

In this section, we prove the main theorem. Let n{\mathbb{Z}}^{n} be the nn-dimensional integer lattice graph, and

ei:=(0,,0i1,1𝑖,0i+1,0),1in,e_{i}:=(0,\cdots,\underset{i-1}{0},\underset{i}{1},\underset{i+1}{0},\cdots 0),\quad 1\leq i\leq n,

be the standard basis of n.{\mathbb{R}}^{n}. We denote by

Q={x=(x1,x2,,xn)n:|xi|12,1in}Q=\{x=(x_{1},x_{2},\cdots,x_{n})\in{\mathbb{R}}^{n}:|x_{i}|\leq\frac{1}{2},\ \forall 1\leq i\leq n\}

the unit cube centered at the origin in n,{\mathbb{R}}^{n}, and by

Qi:=Q{xn:xi=0},1in,Q_{i}:=Q\cap\{x\in{\mathbb{R}}^{n}:x_{i}=0\},\quad 1\leq i\leq n,

the (n1)(n-1)-dimensional unit cube in xix_{i}-hyperplane. For any edge τ={x,y}\tau=\{x,y\} in n,{\mathbb{Z}}^{n}, there is a unique eie_{i} such that yx=eiy-x=e_{i} or ei,-e_{i}, and we define

τ:=12(x+y)+Qi.\tau^{\perp}:=\frac{1}{2}(x+y)+Q_{i}.

That is, τ\tau^{\perp} is an (n1)(n-1)-dimensional unit cube centered at the point 12(x+y)\frac{1}{2}(x+y) parallel to xix_{i}-hyperplane. Note that such a cube τ\tau^{\perp} is one-to-one corresponding to τ.\tau.

For any finite subset Ωn,\Omega\subset{\mathbb{Z}}^{n}, we denote

Ω:={τ:τΩ}.\partial\Omega^{\perp}:=\{\tau^{\perp}:\tau\in\partial\Omega\}.

Note that for any τΩ,\tau\in\partial\Omega, there is a unique end-vertex of the edge τ\tau belonging to δΩ.\delta\Omega. We define a mapping

P:Ω\displaystyle P:\partial\Omega^{\perp} δΩ\displaystyle\rightarrow\delta\Omega
τ\displaystyle\tau^{\perp} P(τ),\displaystyle\mapsto P(\tau^{\perp}),

where P(τ)P(\tau^{\perp}) is the end-vertex of τ\tau in δΩ.\delta\Omega.

Lemma 3.1.

For any τ1\tau_{1},τ2Ω\tau_{2}\in\partial\Omega satisfying P(τ1)P(τ2),P(\tau_{1}^{\perp})\neq P(\tau_{2}^{\perp}), we have

τ1τ2|st|2𝑑n1(s)𝑑n1(t)C3|P(τ1)P(τ2)|2,\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\leq C_{3}\cdot|P(\tau_{1}^{\perp})-P(\tau_{2}^{\perp})|^{2},

where C3=4n,C_{3}=4n, and n1{\mathcal{H}}^{n-1} denotes the (n1)(n-1)-dimensional Hausdorff measure in n.{\mathbb{R}}^{n}.

Proof.

Take τj={xj,yj},\tau_{j}=\{x_{j},y_{j}\}, j=1,2,j=1,2, such that there exists eije_{i_{j}} satisfying yjxj=eij.y_{j}-x_{j}=e_{i_{j}}. By the definition τ1=12(x1+y1)+Qi1.\tau^{\perp}_{1}=\frac{1}{2}(x_{1}+y_{1})+Q_{i_{1}}. By the symmetry, without loss of generality we may assume that P(τ1)=x1.P(\tau_{1}^{\perp})=x_{1}. For any sτ1,s\in\tau_{1}^{\perp}, we write s=12(x1+y1)+qi1s=\frac{1}{2}(x_{1}+y_{1})+q_{i_{1}} where qi1Qi1.q_{i_{1}}\in Q_{i_{1}}. Then,

|sP(τ1)|=|12(x1+y1)+qi1x1|=|ei12+qi1|n2.|s-P(\tau_{1}^{\perp})|=|\frac{1}{2}(x_{1}+y_{1})+q_{i_{1}}-x_{1}|=|\frac{e_{i_{1}}}{2}+q_{i_{1}}|\leq\frac{\sqrt{n}}{2}.

Similarly, we get for any tτ2t\in\tau_{2}^{\perp}

|tP(τ2)|n2.|t-P(\tau_{2}^{\perp})|\leq\frac{\sqrt{n}}{2}.

Since |P(τ1)P(τ2)|1|P(\tau_{1}^{\perp})-P(\tau_{2}^{\perp})|\geq 1, by the triangle inequality, for any sτ1,tτ2s\in\tau_{1}^{\perp},t\in\tau_{2}^{\perp} we have

|st|\displaystyle|s-t| \displaystyle\leq |sP(τ1)|+|P(τ1)P(τ2)|+|P(τ2)t|\displaystyle|s-P(\tau_{1}^{\perp})|+|P(\tau_{1}^{\perp})-P(\tau_{2}^{\perp})|+|P(\tau_{2}^{\perp})-t|
\displaystyle\leq (n+1)|P(τ1)P(τ2)|.\displaystyle(\sqrt{n}+1)|P(\tau_{1}^{\perp})-P(\tau_{2}^{\perp})|.

Hence,

τ1τ2|st|2𝑑n1(s)𝑑n1(t)\displaystyle\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t) \displaystyle\leq (n+1)2|P(τ1)P(τ2)|2\displaystyle(\sqrt{n}+1)^{2}|P(\tau_{1}^{\perp})-P(\tau_{2}^{\perp})|^{2}
\displaystyle\leq 4n|P(τ1)P(τ2)|2.\displaystyle 4n|P(\tau_{1}^{\perp})-P(\tau_{2}^{\perp})|^{2}.

This proves the lemma. ∎

For any r>0,r>0, we denote by

Qr(x):={yn:max1in|yixi|r2}Q_{r}(x):=\{y\in{\mathbb{R}}^{n}:\max_{1\leq i\leq n}|y_{i}-x_{i}|\leq\frac{r}{2}\}

the cube of side length rr centered at x.x. For any subset Ωn,\Omega\subset{\mathbb{Z}}^{n}, we construct a domain Ω^\hat{\Omega} in n{\mathbb{R}}^{n} as follows

Ω^:=xΩQ1(x).\hat{\Omega}:=\bigcup_{x\in\Omega}Q_{1}(x).

Note that the topological boundary of Ω^,\hat{\Omega}, denoted by (Ω^),\partial(\hat{\Omega}), is one-to-one corresponding to Ω,\partial\Omega^{\perp}, i.e.

(3.1) (Ω^)=τΩτ.\partial(\hat{\Omega})=\bigcup_{\tau\in\partial\Omega}\tau^{\perp}.

For our purposes, we will use the geometry of Ω^\hat{\Omega}, which is a domain in n{\mathbb{R}}^{n}, to estimate that of Ω,\Omega, a subset in n.{\mathbb{Z}}^{n}.

We give the definition of “bad” boundary vertices on δΩ.\delta\Omega.

Definition 3.1 (“Bad” boundary vertex).

We call xδΩx\in\delta\Omega a “bad” boundary vertex if it is isolated by the set Ω,\Omega, i.e. for any yxy\sim x in n,{\mathbb{Z}}^{n}, yΩ.y\in\Omega. The set of “bad” boundary vertices is denoted by δΩ.\delta^{\prime}\Omega.

In the following, we give an example to present the role of “bad” boundary vertices in our estimates, which motivates us to distinguish boundary vertices. In the proof of Theorem 1.2, following Brock [Bro01] we reduce the estimate of Steklov eigenvalues to the lower bound estimate of zδΩωδΩ|zω|2,\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}, see (3.8) below. The strategy is to find its counterpart in the continuous setting for deriving such a lower bound, which can be estimated by isoperimetric inequalities in n.{\mathbb{R}}^{n}. A natural choice is the quantity

(Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t)=2n1((Ω^))(Ω^)|sc|2𝑑n1(s),\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)=2{\mathcal{H}}^{n-1}(\partial(\hat{\Omega}))\int_{\partial(\hat{\Omega})}|s-c|^{2}d{\mathcal{H}}^{n-1}(s),

where c=1n1((Ω^))(Ω^)s𝑑n1(s)c=\frac{1}{{\mathcal{H}}^{n-1}(\partial(\hat{\Omega}))}\int_{\partial(\hat{\Omega})}sd{\mathcal{H}}^{n-1}(s) is the barycenter of (Ω^)\partial(\hat{\Omega}) and the right hand side is the derivation of the position vector of (Ω^)\partial(\hat{\Omega}) with respect to the measure n1.{\mathcal{H}}^{n-1}. For the desired estimate

(3.2) zδΩωδΩ|zω|2C(Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t)+,\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}\geq C\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)+\cdots,

we give the following example to indicate the difficulty.

Example 3.2.

For 2,{\mathbb{Z}}^{2}, let

Ω:={(x1,x2)2|max{|x1|,|x2|}2,(x1,x2)(0,0)},\Omega:=\{(x_{1},x_{2})\in{\mathbb{Z}}^{2}|\max\{|x_{1}|,|x_{2}|\}\leq 2,(x_{1},x_{2})\neq(0,0)\},

see Figure 1. Set A1:=δΩ={(0,0)},A2:=δΩδΩ,A_{1}:=\delta^{\prime}\Omega=\{(0,0)\},A_{2}:=\delta\Omega\setminus\delta^{\prime}\Omega, B1:=Q,B_{1}:=\partial Q, the boundary of the unit cube, and B2:=(Q5((0,0))).B_{2}:=\partial(Q_{5}((0,0))). Then (Ω^)=B1B2.\partial(\hat{\Omega})=B_{1}\cup B_{2}. We have the following

zδΩωδΩ|zω|2=(zA1ωA1+2zA1ωA2+zA2ωA2)|zω|2,\displaystyle\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}=\left(\sum_{z\in A_{1}}\sum_{\omega\in A_{1}}+2\sum_{z\in A_{1}}\sum_{\omega\in A_{2}}+\sum_{z\in A_{2}}\sum_{\omega\in A_{2}}\right)|z-\omega|^{2},
(Ω^)(Ω^)|st|2𝑑1(s)𝑑1(t)=(B1B1+2B1B2+B2B2)|st|2d1(s)d1(t).\displaystyle\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{1}(s)d{\mathcal{H}}^{1}(t)=\left(\int_{B_{1}}\int_{B_{1}}+2\int_{B_{1}}\int_{B_{2}}+\int_{B_{2}}\int_{B_{2}}\right)|s-t|^{2}d{\mathcal{H}}^{1}(s)d{\mathcal{H}}^{1}(t).

For deriving the estimate (3.2), it is impossible to bound the term

B1B1|st|2𝑑1(s)𝑑1(t)>0\int_{B_{1}}\int_{B_{1}}|s-t|^{2}d{\mathcal{H}}^{1}(s)d{\mathcal{H}}^{1}(t)>0

from above by the term zA1ωA1|zω|2=0.\sum_{z\in A_{1}}\sum_{\omega\in A_{1}}|z-\omega|^{2}=0. The other terms are mutually comparable. This suggests that “bad” boundary vertices, A1A_{1} here, are the obstructions for the estimate (3.2).

Refer to captionB2B_{2}B1B_{1}A2A_{2}A1A_{1}
Figure 1. Ai,Bi(i=1,2)A_{i},B_{i}~(i=1,2) on Ω.\Omega.
Proposition 3.1.

For any finite subset Ω\Omega in n,{\mathbb{Z}}^{n},

|δΩ||Ω|.|\delta^{\prime}\Omega|\leq|\Omega|.
Proof of Proposition 3.1.

We define a mapping

F:δΩ\displaystyle F:\delta^{\prime}\Omega Ω\displaystyle\rightarrow\Omega
x\displaystyle x F(x):=x+e1.\displaystyle\mapsto F(x):=x+e_{1}.

It is easy to see that FF is injective, which yields |δΩ|=|F(δΩ)||Ω|.|\delta^{\prime}\Omega|=|F(\delta^{\prime}\Omega)|\leq|\Omega|. This proves the proposition. ∎

The next example shows that the above estimate is sharp.

Example 3.3.

For R,R\in{\mathbb{N}}, set

ΩR:={(x,y)2:R=|x||y|orR=|y||x|}{(x,y)2:|x|R1,|y|R1s.t.x+yisodd}.\Omega_{R}:=\{(x,y)\in{\mathbb{Z}}^{2}:R=|x|\geq|y|~\text{or}~R=|y|\geq|x|\}\cup\{(x,y)\in{\mathbb{Z}}^{2}:|x|\leq R-1,~|y|\leq R-1~\mathrm{s.t.}\ x+y\ \mathrm{is\ odd}\}.

Then

limR|δΩR||ΩR|=1.\lim_{R\to\infty}\frac{|\delta^{\prime}\Omega_{R}|}{|\Omega_{R}|}=1.

For any subset Ωn,\Omega\subset{\mathbb{Z}}^{n}, any vertex in VΩ¯V\setminus\bar{\Omega} is called an exterior vertex of Ω.\Omega. For simplicity, we write Ωe:=VΩ¯\Omega^{e}:=V\setminus\bar{\Omega} for the set of exterior vertices of Ω.\Omega. This yields the decomposition

(3.3) n=ΩδΩΩe,{\mathbb{Z}}^{n}=\Omega\sqcup\delta\Omega\sqcup\Omega^{e},

where \sqcup denotes the disjoint union.

Lemma 3.2.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. For any xδΩδΩ,x\in\delta\Omega\setminus\delta^{\prime}\Omega,

|Q2(x)δΩ|2.|Q_{2}(x)\cap\delta\Omega|\geq 2.
Proof.

For any xδΩ,x\in\delta\Omega, set N(x):={yn:yx}.N(x):=\{y\in{\mathbb{Z}}^{n}:y\sim x\}. Without loss of generality, we may assume that N(x)δΩ=,N(x)\cap\delta\Omega=\emptyset, otherwise the lemma follows trivially. Thus, with the decomposition (3.3), we have N(x)ΩΩe.N(x)\subset\Omega\cup\Omega^{e}. Moreover, by the assumption xδΩ,x\not\in\delta^{\prime}\Omega, we get N(x)ΩN(x)\not\subset\Omega so that

(3.4) N(x)Ωe,andN(x)Ω.N(x)\cap\Omega^{e}\neq\emptyset,\ \mathrm{and}\ N(x)\cap\Omega\neq\emptyset.

For any 1i<jn,1\leq i<j\leq n, we set Nij={x±ei,x±ej},N_{ij}=\{x\pm e_{i},x\pm e_{j}\}, which is the intersection of N(x)N(x) with the 22-dimensional affine plane passing through xx spanned by the direction {ei,ej}.\{e_{i},e_{j}\}. We claim that there exist i0i_{0} and j0,j_{0}, 1i0<j0n,1\leq i_{0}<j_{0}\leq n, such that

Ni0j0Ωe,andNi0j0Ω.N_{i_{0}j_{0}}\cap\Omega^{e}\neq\emptyset,\ \mathrm{and}\ N_{i_{0}j_{0}}\cap\Omega\neq\emptyset.

Suppose that it is not true, then for any pair {i,j},\{i,j\}, 1i<jn,1\leq i<j\leq n, NijΩN_{ij}\subset\Omega or NijΩe.N_{ij}\subset\Omega^{e}. Without loss of generality, we assume that N12Ω.N_{12}\subset\Omega. Then x±e1Ω,x\pm e_{1}\in\Omega, which yields that N1jΩN_{1j}\subset\Omega for any j2j\geq 2 by the above condition. This implies that x±ejΩx\pm e_{j}\in\Omega for any 1jn1\leq j\leq n and hence N(x)Ω,N(x)\subset\Omega, which contradicts (3.4). This proves the claim. By the claim, for Ni0j0,N_{i_{0}j_{0}}, one easily verifies that there exists yNi0j0ΩeN(x)Ωey\in N_{i_{0}j_{0}}\cap\Omega^{e}\subset N(x)\cap\Omega^{e} and zNi0j0ΩN(x)Ωz\in N_{i_{0}j_{0}}\cap\Omega\subset N(x)\cap\Omega such that

|yz|=2.|y-z|=\sqrt{2}.

Set w:=y+zx.w:=y+z-x. Then

wQ2(x),yw,zw.w\in Q_{2}(x),\ y\sim w,\ z\sim w.

Note that any path in n{\mathbb{Z}}^{n} connecting a vertex in Ωe\Omega^{e} and a vertex in Ω\Omega contains a vertex in δΩ.\delta\Omega. Applying this to the path ywz,y\sim w\sim z, we get wδΩ.w\in\delta\Omega. This proves the lemma. ∎

For our purposes, we classify pairs of boundary edges, Ω×Ω,\partial\Omega\times\partial\Omega, into “good” pairs and “bad” pairs. The motivation is similar to that for “bad” boundary vertices, see Example 3.2.

Definition 3.4.

A pair of boundary edges (τ1,τ2)Ω×Ω(\tau_{1},\tau_{2})\in\partial\Omega\times\partial\Omega is called “good” if P(τ1)P(τ2)P(\tau^{\perp}_{1})\neq P(\tau^{\perp}_{2}) or P(τ1)=P(τ2)δΩ.P(\tau^{\perp}_{1})=P(\tau^{\perp}_{2})\notin\delta^{\prime}\Omega. Otherwise, it is called “bad”. We denote by (Ω)g2(\partial\Omega)^{2}_{g} the set of “good” pairs of boundary vertices, and by (Ω)b2(\partial\Omega)^{2}_{b} the set of “bad” pairs of boundary vertices.

By the definition,

(3.5) Ω×Ω=(Ω)g2(Ω)b2\partial\Omega\times\partial\Omega=(\partial\Omega)^{2}_{g}\sqcup(\partial\Omega)^{2}_{b}

and

(Ω)b2={(τ1,τ2)Ω×Ω|P(τ1)=P(τ2)δΩ}.(\partial\Omega)^{2}_{b}=\{(\tau_{1},\tau_{2})\in\partial\Omega\times\partial\Omega~|P(\tau^{\perp}_{1})=P(\tau^{\perp}_{2})\in\delta^{\prime}\Omega\}.
Definition 3.5 (The multiplicity of a mapping).

Let T:ABT:A\to B be a mapping. The multiplicity of the mapping TT is defined as

mT:=supbB|{T1(b)}|.m_{T}:=\sup_{b\in B}|\{T^{-1}(b)\}|.

We estimate the number of “bad” pairs of boundary vertices by the number of “bad” vertices.

Proposition 3.2.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. Then

|(Ω)b2|4n2|δΩ|.|(\partial\Omega)^{2}_{b}|\leq 4n^{2}|\delta^{\prime}\Omega|.
Proof.

We define a mapping

H:(Ω)b2\displaystyle H:(\partial\Omega)^{2}_{b} δΩ\displaystyle\rightarrow\delta^{\prime}\Omega
(τ1,τ2)\displaystyle(\tau_{1},\tau_{2}) H(τ1,τ2)=P(τ1).\displaystyle\mapsto H(\tau_{1},\tau_{2})=P(\tau_{1}^{\perp}).

We estimate the multiplicity of the mapping H.H. Fix any xδΩ.x\in\delta^{\prime}\Omega. For any (τ1,τ2)(Ω)b2(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{b} such that H(τ1,τ2)=x,H(\tau_{1},\tau_{2})=x,

P(τ1)=P(τ2)=x.P(\tau_{1}^{\perp})=P(\tau_{2}^{\perp})=x.

So that xx is a common end-vertex of τ1\tau_{1} and τ2.\tau_{2}. The number of possible pairs (τ1,τ2)(\tau_{1},\tau_{2}) satisfying this property is at most 4n2.4n^{2}. Hence mT4n2.m_{T}\leq 4n^{2}. This proves the proposition. ∎

We define a mapping on“good” pairs of boundary vertices as

(3.6) f:(Ω)g2\displaystyle f:(\partial\Omega)^{2}_{g} δΩ×δΩ\displaystyle\rightarrow\delta\Omega\times\delta\Omega
(τ1,τ2)\displaystyle(\tau_{1},\tau_{2}) f(τ1,τ2)=(f1(τ1,τ2),f2(τ1,τ2)):={(P(τ1),P(τ2)),if P(τ1)P(τ2)(P(τ1),z),if P(τ1)=P(τ2)δΩ,\displaystyle\mapsto f(\tau_{1},\tau_{2})=(f_{1}(\tau_{1},\tau_{2}),f_{2}(\tau_{1},\tau_{2})):=\begin{cases}(P(\tau_{1}^{\perp}),P(\tau_{2}^{\perp})),&\text{if }P(\tau_{1}^{\perp})\neq P(\tau_{2}^{\perp})\\ (P(\tau_{1}^{\perp}),z),&\text{if }P(\tau_{1}^{\perp})=P(\tau_{2}^{\perp})\not\in\delta^{\prime}\Omega,\end{cases}

where zQ2(P(τ1))δΩz\in Q_{2}(P(\tau_{1}^{\perp}))\cap\delta\Omega such that zP(τ1).z\neq P(\tau_{1}^{\perp}).

Remark 3.6.

In the definition of the mapping f,f, the existence of zQ2(P(τ1))δΩz\in Q_{2}(P(\tau_{1}^{\perp}))\cap\delta\Omega in the second case, which is not necessarily unique, is guaranteed by Lemma 3.2. Any choice of such a vertex zz will be sufficient for our purposes.

Lemma 3.3.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. Then for any (τ1,τ2)(Ω)g2,(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{g},

τ1τ2|st|2𝑑n1(s)𝑑n1(t)C3|f1(τ1,τ2)f2(τ1,τ2)|2,\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\leq C_{3}\cdot|f_{1}(\tau_{1},\tau_{2})-f_{2}(\tau_{1},\tau_{2})|^{2},

where C3C_{3} is the constant in Lemma 3.1.

Proof of Lemma 3.3.

By Lemma 3.1, it suffices to consider the case that P(τ1)=P(τ2)δΩ.P(\tau_{1}^{\perp})=P(\tau_{2}^{\perp})\not\in\delta^{\prime}\Omega. Set X:=P(τ1)=P(τ2).X:=P(\tau_{1}^{\perp})=P(\tau_{2}^{\perp}). Then τ1τ2=X.\tau_{1}\cap\tau_{2}=X. By the triangle inequality, for any sτ1,s\in\tau_{1}^{\perp}, tτ2,t\in\tau_{2}^{\perp}, we have

(3.7) |st||sX|+|Xt|n.|s-t|\leq|s-X|+|X-t|\leq\sqrt{n}.

Then τ1τ2|st|2𝑑n1(s)𝑑n1(t)n.\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\leq n. Obviously, since |f1(τ1,τ2)f2(τ1,τ2)|=|P(τ1)z|1,|f_{1}(\tau_{1},\tau_{2})-f_{2}(\tau_{1},\tau_{2})|=|P(\tau_{1}^{\perp})-z|\geq 1,

τ1τ2|st|2𝑑n1(s)𝑑n1(t)n|f1(τ1,τ2)f2(τ1,τ2)|2C3|f1(τ1,τ2)f2(τ1,τ2)|2.\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\leq n\cdot|f_{1}(\tau_{1},\tau_{2})-f_{2}(\tau_{1},\tau_{2})|^{2}\leq C_{3}\cdot|f_{1}(\tau_{1},\tau_{2})-f_{2}(\tau_{1},\tau_{2})|^{2}.

This proves the lemma. ∎

By the definition of the multiplicity of f,f, we have the following proposition.

Proposition 3.3.
(τ1,τ2)(Ω)g2|f1(τ1,τ2)f2(τ1,τ2)|2mfzδΩωδΩ|zω|2.\sum_{(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{g}}|f_{1}(\tau_{1},\tau_{2})-f_{2}(\tau_{1},\tau_{2})|^{2}\leq m_{f}\cdot\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}.
Lemma 3.4.

Let Ω\Omega be a finite subset of n.{\mathbb{Z}}^{n}. Then

mf8n2.m_{f}\leq 8n^{2}.
Proof.

Fix (x,y)f((Ω)g2).(x,y)\in f((\partial\Omega)^{2}_{g}). Let (τ1,τ2)(Ω)g2(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{g} such that f(τ1,τ2)=(x,y).f(\tau_{1},\tau_{2})=(x,y). We estimate the number of possible pairs (τ1,τ2)(\tau_{1},\tau_{2}) satisfying the above property. By the definition of ff in (3.6), f1(τ1,τ2)=P(τ1)=x,f_{1}(\tau_{1},\tau_{2})=P(\tau_{1}^{\perp})=x, which yields that xx is an end-vertex of the edge τ1.\tau_{1}. Hence, the number of possible choices of τ1\tau_{1} is at most 2n.2n. To determine the edge τ2,\tau_{2}, it is divided into the following cases.

  1. Case 1.

    P(τ1)P(τ2).P(\tau_{1}^{\perp})\neq P(\tau_{2}^{\perp}). By (3.6),

    f2(τ1,τ2)=P(τ2)=y.f_{2}(\tau_{1},\tau_{2})=P(\tau_{2}^{\perp})=y.

    Hence, yy is an end-vertex of the edge τ2.\tau_{2}. So that the number of possible choices of τ2\tau_{2} is at most 2n.2n.

  2. Case 2.

    P(τ1)=P(τ2)δΩ.P(\tau_{1}^{\perp})=P(\tau_{2}^{\perp})\not\in\delta^{\prime}\Omega. In this case, P(τ2)=xP(\tau_{2}^{\perp})=x and xx is an end-vertex of the edge τ2.\tau_{2}. Hence, the number of possible choices of τ2\tau_{2} is at most 2n.2n.

Combining the above cases, the number of possible choices of τ2\tau_{2} is at most 4n.4n. This yields that

mf2n×4n=8n2.m_{f}\leq 2n\times 4n=8n^{2}.

This proves the lemma. ∎

The following lemma is a version of weighted isoperimetric inequality by symmetrization approaches, see e.g. [BBMP99], which is crucial in the proof of Brock’s result in n.{\mathbb{R}}^{n}. For any R>0,R>0, we denote by BRB_{R} the ball of radius RR centered at the origin in n.{\mathbb{R}}^{n}.

Lemma 3.5.

Let UU be a bounded domain in n{\mathbb{R}}^{n} with Lipschitz boundary. Let R>0R>0 such that n(BR)=n(U).{\mathcal{H}}^{n}(B_{R})={\mathcal{H}}^{n}(U). Let gC([0,+))g\in C([0,+\infty)) be nonnegative, nondecreasing and suppose that

(g(z1n)g(0))z11n,z0(g(z^{\frac{1}{n}})-g(0))z^{1-\frac{1}{n}},~z\geq 0

is convex. Then

Ug(|x|)𝑑SBRg(|x|)𝑑S=nωng(R)Rn1.\int_{\partial U}g(|x|)dS\geq\int_{\partial B_{R}}g(|x|)dS=n\omega_{n}g(R)R^{n-1}.

Now we are ready to prove our main result.

Proof of Theorem 1.2.

For 1kn,1\leq k\leq n, we denote by EkE_{k} the edges in E(Ω,Ω¯)E(\Omega,\bar{\Omega}) parallel to ek,e_{k}, and by |Ek||E_{k}| the cardinality of Ek.E_{k}. Using the coordinate functions of n{\mathbb{R}}^{n}, we define functions uinu_{i}\in{\mathbb{R}}^{{\mathbb{Z}}^{n}} for 2in+12\leq i\leq n+1 as follows,

ui(z):=|Ei1|12(zi1zi1¯),u_{i}(z):=|E_{i-1}|^{-\frac{1}{2}}(z_{i-1}-\overline{z_{i-1}}),

where

zi1¯:=1|δΩ|zδΩzi1.\overline{z_{i-1}}:=\frac{1}{|\delta\Omega|}\sum_{z\in\delta\Omega}z_{i-1}.

Then, we get

DΩ(ui,uj)=δij,zδΩui(z)=0,2i,jn+1.D_{\Omega}(u_{i},u_{j})=\delta_{ij},\ \sum_{z\in\delta\Omega}u_{i}(z)=0,\quad 2\leq i,j\leq n+1.

Hence, by the variational principle, Theorem 2.2 with p=n+1p=n+1,

i=2n+11λi(Ω)\displaystyle\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)} \displaystyle\geq i=2n+1zδΩui2(z)=i=2n+1zδΩ|zi1zi1¯|2|Ei1|\displaystyle\sum^{n+1}_{i=2}\sum_{z\in\delta\Omega}u^{2}_{i}(z)=\sum^{n+1}_{i=2}\sum_{z\in\delta\Omega}\frac{|z_{i-1}-\overline{z_{i-1}}|^{2}}{|E_{i-1}|}
\displaystyle\geq 1|E(Ω,Ω¯)|i=2n+1zδΩ|zi1zi1¯|2.\displaystyle\frac{1}{|E(\Omega,\bar{\Omega})|}\sum^{n+1}_{i=2}\sum_{z\in\delta\Omega}|z_{i-1}-\overline{z_{i-1}}|^{2}.

Note that

zδΩωδΩ|zω|2\displaystyle\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}
=\displaystyle= zδΩωδΩi=2n+1|zi1ωi1|2=i=2n+1(2|δΩ|zδΩzi122(zδΩzi1)2)\displaystyle\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}\sum^{n+1}_{i=2}|z_{i-1}-\omega_{i-1}|^{2}=\sum^{n+1}_{i=2}\left(2|\delta\Omega|\sum_{z\in\delta\Omega}z^{2}_{i-1}-2(\sum_{z\in\delta\Omega}z_{i-1})^{2}\right)
=\displaystyle= 2|δΩ|i=2n+1(zδΩzi12(zδΩzi1)2|δΩ|)=2|δΩ|zδΩi=2n+1(zi122zi1¯zi1+zi1¯2)\displaystyle 2|\delta\Omega|\sum^{n+1}_{i=2}\left(\sum_{z\in\delta\Omega}z^{2}_{i-1}-\frac{(\sum_{z\in\delta\Omega}z_{i-1})^{2}}{|\delta\Omega|}\right)=2|\delta\Omega|\sum_{z\in\delta\Omega}\sum^{n+1}_{i=2}(z^{2}_{i-1}-2\overline{z_{i-1}}\cdot z_{i-1}+\overline{z_{i-1}}^{2})
=\displaystyle= 2|δΩ|zδΩi=2n+1(zi1zi1¯)2.\displaystyle 2|\delta\Omega|\sum_{z\in\delta\Omega}\sum^{n+1}_{i=2}(z_{i-1}-\overline{z_{i-1}})^{2}.

Hence, we get

(3.8) i=2n+11λi(Ω)12|δΩ|1|E(Ω,Ω¯)|zδΩωδΩ|zω|2.\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)}\geq\frac{1}{2|\delta\Omega|}\cdot\frac{1}{|E(\Omega,\bar{\Omega})|}\cdot\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}.

In the following, we compare the right hand side of the above inequality, a discrete summation, with a continuous quantity,

(Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t).\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t).

By (3.1) and the decomposition of pairs of boundary edges into “good” and “bad” ones, (3.5), we have

(Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t)\displaystyle\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t) =(τ1,τ2)Ω×Ωτ1τ2|st|2𝑑n1(s)𝑑n1(t)\displaystyle=\sum_{(\tau_{1},\tau_{2})\in\partial\Omega\times\partial\Omega}\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)
=((τ1,τ2)(Ω)g2+(τ1,τ2)(Ω)b2)τ1τ2|st|2𝑑n1(s)𝑑n1(t)\displaystyle=\left(\sum_{(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{g}}+\sum_{(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{b}}\right)\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)
=:I+II.\displaystyle=:I+II.

By Lemma 3.3 and Proposition 3.3,

I\displaystyle I =\displaystyle= (τ1,τ2)(Ω)g2τ1τ2|st|2𝑑n1(s)𝑑n1(t)C3mfzδΩωδΩ|zω|2.\displaystyle\sum_{(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{g}}\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\leq C_{3}\cdot m_{f}\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}.

For any (τ1,τ2)(Ω)b2,(\tau_{1},\tau_{2})\in(\partial\Omega)^{2}_{b}, the inequality (3.7) yields

τ1τ2|st|2𝑑n1(s)𝑑n1(t)n.\int_{\tau_{1}^{\perp}}\int_{\tau_{2}^{\perp}}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\leq n.

Combining this with Proposition 3.2, we get

II4n3|δΩ|.II\leq 4n^{3}|\delta^{\prime}\Omega|.

Hence, we obtain

(3.9) (Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t)=I+IIC3mfzδΩωδΩ|zω|2+4n3|δΩ|.\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)=I+II\leq C_{3}\cdot m_{f}\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}+4n^{3}|\delta^{\prime}\Omega|.

By the calculation,

(3.10) (Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t)=2n1((Ω^))(Ω^)|sc|2𝑑n1(s),\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)=2{\mathcal{H}}^{n-1}(\partial(\hat{\Omega}))\int_{\partial(\hat{\Omega})}|s-c|^{2}d{\mathcal{H}}^{n-1}(s),

where c=1n1((Ω^))(Ω^)s𝑑n1(s)c=\frac{1}{{\mathcal{H}}^{n-1}(\partial(\hat{\Omega}))}\int_{\partial(\hat{\Omega})}sd{\mathcal{H}}^{n-1}(s) is the barycenter of (Ω^).\partial(\hat{\Omega}). We denote by Ω0:=Ω^c\Omega_{0}:=\hat{\Omega}-c the translation of the domain Ω^\hat{\Omega} by a vector c.-c. It is easy to see that the barycenter of the boundary of Ω0\Omega_{0} is the origin, so that

(Ω^)|sc|2𝑑n1(s)=Ω0|s|2𝑑n1(s).\int_{\partial(\hat{\Omega})}|s-c|^{2}d{\mathcal{H}}^{n-1}(s)=\int_{\partial\Omega_{0}}|s|^{2}d{\mathcal{H}}^{n-1}(s).

Since n(Ω0)=n(Ω^)=|Ω|,{\mathcal{H}}^{n}(\Omega_{0})={\mathcal{H}}^{n}(\hat{\Omega})=|\Omega|, we choose R>0,R>0, such that

n(BR)=ωnRn=|Ω|=n(Ω0).{\mathcal{H}}^{n}(B_{R})=\omega_{n}R^{n}=|\Omega|={\mathcal{H}}^{n}(\Omega_{0}).

By Lemma 3.5, taking U=Ω0U=\Omega_{0} and g(t)=t2,g(t)=t^{2}, we get

Ω0|s|2𝑑n1(s)BR|s|2𝑑n1(s)=nωnRn+1=n|Ω|R.\int_{\partial\Omega_{0}}|s|^{2}d{\mathcal{H}}^{n-1}(s)\geq\int_{\partial B_{R}}|s|^{2}d{\mathcal{H}}^{n-1}(s)=n\omega_{n}R^{n+1}=n|\Omega|R.

By the above estimates, noting that n1((Ω^))=|Ω|,{\mathcal{H}}^{n-1}(\partial(\hat{\Omega}))=|\partial\Omega|, we obtain

(3.11) (Ω^)(Ω^)|st|2𝑑n1(s)𝑑n1(t)2n|Ω||Ω|R.\int_{\partial(\hat{\Omega})}\int_{\partial(\hat{\Omega})}|s-t|^{2}d{\mathcal{H}}^{n-1}(s)d{\mathcal{H}}^{n-1}(t)\geq 2n|\partial\Omega|\cdot|\Omega|R.

By the equations (3.9) and (3.11),

(3.12) zδΩωδΩ|zω|22n(|Ω||Ω|R2n2|δΩ|)C3mf.\sum_{z\in\delta\Omega}\sum_{\omega\in\delta\Omega}|z-\omega|^{2}\geq\frac{2n(|\partial\Omega|\cdot|\Omega|R-2n^{2}|\delta^{\prime}\Omega|)}{C_{3}\cdot m_{f}}.

Note that

|δΩ||δΩ||Ω|2n|Ω|,|\delta^{\prime}\Omega|\leq|\delta\Omega|\leq|\partial\Omega|\leq 2n|\Omega|,

and

2|E(Ω,Ω¯)|=2n|Ω|+|Ω|.2|E(\Omega,\bar{\Omega})|=2n\cdot|\Omega|+|\partial\Omega|.

This yields that

|E(Ω,Ω¯)|=n|Ω|+12|Ω|2n|Ω|.|E(\Omega,\bar{\Omega})|=n|\Omega|+\frac{1}{2}|\partial\Omega|\leq 2n|\Omega|.

By (3.8) and (3.12),

i=2n+11λi(Ω)\displaystyle\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)} \displaystyle\geq 2n(|Ω||Ω|R2n2|δΩ|)2|δΩ||E(Ω,Ω¯)|C3mf.\displaystyle\frac{2n(|\partial\Omega|\cdot|\Omega|R-2n^{2}|\delta^{\prime}\Omega|)}{2|\delta\Omega|\cdot|E(\Omega,\bar{\Omega})|C_{3}\cdot m_{f}}.

This yields that

i=2n+11λi(Ω)\displaystyle\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)} \displaystyle\geq n(|Ω||Ω|R2n2|δΩ|)|Ω|2n|Ω|C3mf\displaystyle\frac{n(|\partial\Omega|\cdot|\Omega|R-2n^{2}|\delta^{\prime}\Omega|)}{|\partial\Omega|\cdot 2n|\Omega|C_{3}\cdot m_{f}}
\displaystyle\geq 12C3mf(R2n2|Ω|)=12C3mf((ωn1|Ω|)1n2n2|Ω|).\displaystyle\frac{1}{2C_{3}\cdot m_{f}}\left(R-\frac{2n^{2}}{|\Omega|}\right)=\frac{1}{2C_{3}\cdot m_{f}}\left((\omega_{n}^{-1}|\Omega|)^{\frac{1}{n}}-\frac{2n^{2}}{|\Omega|}\right).

By Lemma 3.4, we get

i=2n+11λi(Ω)C1|Ω|1nC2|Ω|,\sum^{n+1}_{i=2}\frac{1}{\lambda_{i}(\Omega)}\geq C_{1}|\Omega|^{\frac{1}{n}}-\frac{C_{2}}{|\Omega|},

where C1=(64n3ωn1n)1,C2=132n.C_{1}=(64n^{3}\omega_{n}^{\frac{1}{n}})^{-1},C_{2}=\frac{1}{32n}. This proves the theorem. ∎

We construct an infinite graph with bounded degree, which doesn’t satisfy the property (1.6). For any n,n\in{\mathbb{N}}, we denote by TnT_{n} the graph obtained by a complete binary tree of depth nn with a pending vertex PTnP_{T_{n}} attaching to the root of the tree, see Figure 2. Let TnT_{n}^{\prime} be a copy of Tn.T_{n}. Let KnK_{n} be the graph constructed by the disjoint union of TnT_{n} and TnT_{n}^{\prime} by identifying the leaves of TnT_{n} with those of TnT_{n}^{\prime} accordingly. Note that KnK_{n} has two pending vertices PTnP_{T_{n}} and PTn,P_{T_{n}^{\prime}}, see Figure 2.

Refer to captionT2T_{2}K2K_{2}PT2P_{T_{2}}PT2P_{T_{2}^{\prime}}
Figure 2. K2K_{2} is constructed by T2T_{2} and its copy T2T_{2}^{\prime}.
Refer to captionK2K_{2}K3K_{3}
Figure 3. An infinite graph doesn’t satisfy the property (1.6).
Example 3.7.

Let G=(K1K2K3)/G=(K_{1}\cup K_{2}\cup K_{3}\cup\cdots)/\sim be an infinite graph obtained by a disjoint union of {Kn}n=1\{K_{n}\}_{n=1}^{\infty} with sequentially identifying pending vertices, PTnP_{T_{n}^{\prime}} and PTn+1P_{T_{n+1}}, of KnK_{n} and Kn+1,K_{n+1}, n1.n\geq 1. Let Ωi=Ki,\Omega_{i}=K_{i}, i1,i\geq 1, where KiK_{i} is the embedded image of KiK_{i} in G.G. Then

λ2(Ωi)=2Res(Ki¯)=2612i1,i2,\lambda_{2}(\Omega_{i})=\frac{2}{\mathrm{Res}(\overline{K_{i}})}=\frac{2}{6-\frac{1}{2^{i-1}}},\quad\forall i\geq 2,

where Res(Ki¯)\mathrm{Res}(\overline{K_{i}}) is the effective resistance between the vertices in δKi\delta K_{i} in the induced subgraph on Ki¯,\overline{K_{i}}, for which each edge is interpreted as a resistor of resistance one; see [Bar17, Definition 2.7, p.42] for the definition. This yields that

λ2(Ωi)c>0,i1.\lambda_{2}(\Omega_{i})\geq c>0,\quad i\geq 1.

Acknowledgements. We thank the anonymous referees for providing helpful comments and suggestions to improve the writing of the paper, in particular pointing out some useful references on Steklov eigenvalues. We thank Zuoqin Wang for many helpful suggestions on the Steklov problem on graphs.

B. H. is supported by NSFC (China), grant no. 11831004 and no. 11826031.

References

  • [Ban80] C. Bandle. Isoperimetric inequalities and applications, volume 7 of Monographs and Studies in Mathematics. Pitman (Advanced Publishing Program), Boston, Mass.-London, 1980.
  • [Bar17] Martin T. Barlow. Random walks and heat kernels on graphs, volume 438 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2017.
  • [BBMP99] M. F. Betta, F. Brock, A. Mercaldo, and M. R. Posteraro. A weighted isoperimetric inequality and applications to symmetrization. J. Inequal. Appl., 4(3):215–240, 1999.
  • [BFNT17] D. Bucur, V. Ferone, C. Nitsch, and C. Trombetti. Weinstock inequality in higher dimensions. arXiv:1710.04587, 2017.
  • [Bro01] F. Brock. An isoperimetric inequality for eigenvalues of the Stekloff problem. ZAMM Z. Angew. Math. Mech., 81(1):69–71, 2001.
  • [Cal06] A. P. Calderón. On an inverse boundary value problem. Comput. Appl. Math., 25(2-3):133–138, 2006.
  • [CSG11] B. Colbois, A. E. Soufi, and A. Girouard. Isoperimetric control of the Steklov spectrum. J. Funct. Anal., 261(5):1384–1399, 2011.
  • [Esc97] J. Escobar. The geometry of the first non-zero Stekloff eigenvalue. J. Funct. Anal., 150(2):544–556, 1997.
  • [Esc99] J. Escobar. An isoperimetric inequality and the first Steklov eigenvalue. J. Funct. Anal., 165(1):101–116, 1999.
  • [FL14] A. Fraser and M. C. Li. Compactness of the space of embedded minimal surfaces with free boundary in three-manifolds with nonnegative Ricci curvature and convex boundary. J. Differential Geom., 96(2):183–200, 2014.
  • [FS16] A. Fraser and R. Schoen. Sharp eigenvalue bounds and minimal surfaces in the ball. Invent. Math., 203(3):823–890, 2016.
  • [FS19] A. Fraser and R. Schoen. Shape optimization for the Steklov problem in higher dimensions. Adv. Math., 348:146–162, 2019.
  • [FW17] M. M. Fall and T. Weth. Profile expansion for the first nontrivial Steklov eigenvalue in Riemannian manifolds. Comm. Anal. Geom., 25(2):431–463, 2017.
  • [GLS16] A. Girouard, R. S. Laugesen, and B. A. Siudeja. Steklov eigenvalues and quasiconformal maps of simply connected planar domains. Arch. Ration. Mech. Anal., 219(2):903–936, 2016.
  • [GP10] A. Girouard and I. Polterovich. Shape optimization for low Neumann and Steklov eigenvalues. Math. Methods Appl. Sci., 33(4):501–516, 2010.
  • [GP12] A. Girouard and I. Polterovich. Upper bounds for Steklov eigenvalues on surfaces. Electron. Res. Announc. Math. Sci., 19:77–85, 2012.
  • [Has11] A. Hassannezhad. Conformal upper bounds for the eigenvalues of the Laplacian and Steklov problem. J. Funct. Anal., 261(12):3419–3436, 2011.
  • [HHW17] B. Hua, Y. Huang, and Z. Wang. First eigenvalue estimates of Dirichlet-to-Neumann operators on graphs. Calc. Var. Partial Differential Equations, 56(6):Art. 178, 21, 2017.
  • [HHW18] B. Hua, Y. Huang, and Z. Wang. Cheeger estimates of Dirichlet-to-Neumann operators on infinite subgraphs of graphs. arXiv:1810.10763, 2018.
  • [HM17] A. Hassannezhad and L. Miclo. Higher order Cheeger inequalities for Steklov eigenvalues. arXiv:1705.08643, to appear in Ann. Scient. Éc. Norm. Sup., doi:10.24033/asens.2417, 2017.
  • [HP68] J. Hersch and L. E. Payne. Extremal principles and isoperimetric inequalities for some mixed problems of Stekloff’s type. Z. Angew. Math. Phys., 19:802–817, 1968.
  • [HPS08] A. Henrot, G. A. Philippin, and A. Safoui. Some isoperimetric inequalities with application to the Stekloff problem. J. Convex Anal., 15(3):581–592, 2008.
  • [KKP14] M. Karpukhin, G. Kokarev, and I. Polterovich. Multiplicity bounds for Steklov eigenvalues on Riemannian surfaces. Ann. Inst. Fourier (Grenoble), 64(6):2481–2502, 2014.
  • [Mat17] H. A. Matevossian. On the biharmonic Steklov problem in weighted spaces. Russ. J. Math. Phys., 24(1):134–138, 2017.
  • [Per19] H. Perrin. Lower bounds for the first eigenvalue of the Steklov problem on graphs. Calc. Var. Partial Differential Equations, 58(2):Art. 67, 12, 2019.
  • [Per20] H. Perrin. Isoperimetric upper bound for the first eigenvalue of discrete steklov problems. arXiv:2002.08751, 2020.
  • [SU90] J. Sylvester and G. Uhlmann. The Dirichlet to Neumann map and applications. In Inverse problems in partial differential equations (Arcata, CA, 1989), pages 101–139. SIAM, Philadelphia, PA, 1990.
  • [Wei54] R. Weinstock. Inequalities for a classical eigenvalue problem. J. Ration. Mech. Anal., 3:745–753, 1954.
  • [WX09] Q. Wang and C. Xia. Sharp bounds for the first non-zero Stekloff eigenvalues. J. Funct. Anal., 257(8):2635–2644, 2009.
  • [YY17] L. Yang and C. Yu. Estimates for higher Steklov eigenvalues. J. Math. Phys., 58(2):021504, 9, 2017.