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

Spectral gap estimates for Brownian motion on domains with sticky-reflecting boundary diffusion

Vitalii Konarovskyi∗†, Victor Marx, and Max von Renesse
Abstract

Introducing an interpolation method we estimate the spectral gap for Brownian motion on general domains with sticky-reflecting boundary diffusion associated to the first nontrivial eigenvalue for the Laplace operator with corresponding Wentzell-type boundary condition. In the manifold case our proofs involve novel applications of the celebrated Reilly formula.

* Universität Leipzig, Fakultät für Mathematik und Informatik, Augustusplatz 10, 04109 Leipzig, Germany; \dagger Universität Hamburg, Bundesstraße 55, 20146 Hamburg, Germany; Institute of Mathematics of NAS of Ukraine, Tereschenkivska st. 3, 01024 Kiev, Ukraine konarovskyi@gmail.com, marx@math.uni-leipzig.de, renesse@uni-leipzig.de Mathematics Subject Classification (2020): Primary 26D10, 35A23, 34K08 ; Secondary 46E35, 53B25, 60J60, 47D07.

1 Introduction and statement of main results

Brownian motion on smooth domains with sticky-reflecting diffusion along the boundary has a long history, dating back at least to Wentzell [34]. As a prototype consider a diffusion on the closure Ω¯\overline{\Omega} of a smooth domain Ω\Omega with Feller generator (𝒟(A),A)(\mathcal{D}(A),A)

𝒟(A)={fC0(Ω¯)|AfC0(Ω¯)}\displaystyle\mathcal{D}(A)=\{f\in C_{0}(\overline{\Omega})\,|\,Af\in C_{0}(\overline{\Omega})\} (1.1)
Af=Δf𝕀Ω+(βΔτfγfν)𝕀Ω\displaystyle Af=\Delta f\mathbb{I}_{\Omega}+(\beta\Delta^{\tau}f-\gamma\frac{\partial f}{\partial\nu})\mathbb{I}_{\partial\Omega}

where ν\frac{\partial}{\partial\nu} is the outer normal derivative, Δτ\Delta^{\tau} is the Laplace-Beltrami operator on the boundary Ω\partial\Omega and β>0,γ\beta>0,\gamma\in\mathbb{R}. The case of pure sticky reflection but no diffusion along the boundary corresponds to the regime β=0\beta=0; models with β>0\beta>0 have appeared recently in interacting particle systems with singular boundary or zero-range pair interaction [1, 7, 13, 19, 27]. The first rigorous process constructions on special domains Ω\Omega were given in [16, 33, 37] and were later extended to jump-diffusion processes on general domains [6] cf. [32]. An efficient construction in symmetric cases was given by Grothaus and Voßhall via Dirichlet forms in [15]. Qualitative regularity properties of the associated semigroups were studied e.g. in [14]. In this note we address the problem of estimating the spectral gap for such processes, which is a natural question also in algorithmic applications. To our knowledge this question has been considered only for β=0\beta=0 by Kennedy [17] and Shouman [30]. However, for β>0\beta>0 the properties of the process change significantly, which is indicated by the fact that the energy form of AA now also contains a boundary part and which also constitutes the main difference to the closely related work [18].

In the sequel we treat the case when γ>0\gamma>0 which corresponds to an inward sticky reflection at Ω\partial\Omega. Our ansatz to estimate the spectral gap is based on a simple interpolation idea. To this aim assume that Ω\Omega and Ω\partial\Omega have finite (Hausdorff) measure so that we may choose α(0,1)\alpha\in(0,1) for which

α1α|Ω||Ω|=γ.\frac{\alpha}{1-\alpha}\frac{|\partial\Omega|}{|\Omega|}=\gamma.

Introducing λΩ\lambda_{\Omega} and λ\lambda_{\partial} as normalized volume and Hausdorff measures on Ω\Omega and Ω\partial\Omega and setting

λα=αλΩ+(1α)λ,\lambda_{\alpha}=\alpha\lambda_{\Omega}+(1-\alpha)\lambda_{\partial},

we find that A-A is λα\lambda_{\alpha}-symmetric with first nonzero eigenvalue/spectral gap characterized by the Rayleigh quotient

σα,β=inffC1(Ω¯)Varλα(f)>0α,β(f)Varλαf,\sigma_{\alpha,\beta}=\inf_{\begin{subarray}{c}f\in C^{1}(\overline{\Omega})\\ \operatorname{Var}_{\lambda_{\alpha}}(f)>0\end{subarray}}\frac{\mathcal{E}_{\alpha,\beta}(f)}{\operatorname{Var}_{\lambda_{\alpha}}f},

where

Varλαf=Ωf2𝑑λα(Ωf𝑑λα)2\operatorname{Var}_{\lambda_{\alpha}}f=\int_{\Omega}f^{2}d\lambda_{\alpha}-\left(\int_{\Omega}fd\lambda_{\alpha}\right)^{2}

and

α,β(f)=αΩf2𝑑λΩ+(1α)Ωβτf2𝑑λ,\mathcal{E}_{\alpha,\beta}(f)=\alpha\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}+(1-\alpha)\int_{\partial\Omega}\beta\|\nabla^{\tau}f\|^{2}d\lambda_{\partial},

and τ\nabla^{\tau} denotes the tangential derivative operator on Ω\partial\Omega.

This representation of σα,β\sigma_{\alpha,\beta} formally interpolates between the two extremal cases of the spectral gap for reflecting Brownian motion on Ω\Omega when α=1\alpha=1 and for Brownian motion on the surface Ω\partial\Omega when α=0\alpha=0. As our main result, in Proposition 2.1 we propose a simple method to estimate σα,β\sigma_{\alpha,\beta} from below using only σ0\sigma_{0} and σ1\sigma_{1} and estimates for certain bulk-boundary interaction terms which are independent of α\alpha. The method can lead to quite good results which is illustrated by the example when Ω=B1d\Omega=B_{1}\subset\mathbb{R}^{d} is a dd-dimensional unit ball. When d=2d=2 and β=1\beta=1, for instance, it yields the estimate

σα8(1+α)σΩ8(1α)σΩ+16α+3α(1α)σΩ with α=γ2+γ,\displaystyle\sigma_{\alpha}\geq\frac{8(1+\alpha)\sigma_{\Omega}}{8(1-\alpha)\sigma_{\Omega}+16\alpha+3\alpha(1-\alpha)\sigma_{\Omega}}\mbox{ with }\alpha=\frac{\gamma}{2+\gamma},

where σ03.39\sigma_{0}\approx 3.39 is the spectral gap for the Neumann Laplacian on the 2-dimensional unit ball, c.f.  Section 3.1. – In case when Ω\Omega is a dd-dimensional manifold with Ricci curvature bounded from below by kR>0k_{R}>0 and with boundary Ω\partial\Omega whose second fundamental form IIΩ\mathop{\rm II}_{\partial\Omega} is bounded from below by k2>0k_{2}>0 we obtain (again with β=1\beta=1, for simplicity) that

σαmin(dkRCΩdkR+(1α)(d1),dkRCΩ2(1α)+αk2CΩ2(1α)dkR+αdk2kRCΩ+α(1α)(d1)k2),\displaystyle\sigma_{\alpha}\geq\min\left(\frac{dk_{R}}{C_{\Omega}dk_{R}+(1-\alpha)(d-1)},\frac{dk_{R}}{C_{\partial\Omega}}\frac{2(1-\alpha)+\alpha k_{2}C_{\partial\Omega}}{2(1-\alpha)dk_{R}+\alpha dk_{2}k_{R}C_{\Omega}+\alpha(1-\alpha)(d-1)k_{2}}\right),

where CΩC_{\Omega} and CΩC_{\partial\Omega} are the usual (Neumann) Poincaré constants of Ω\Omega and Ω\partial\Omega respectively. To derive this result we combine Escobar’s lower bound [9] on the first Steklov eigenvalue [12, 20] of Ω\Omega with a novel estimate on the optimal zero mean trace Poincaré constant of Ω\Omega [22, 26], for which we obtain that

Ωf2𝑑xd1dkRΩ|f|2,\int_{\Omega}f^{2}dx\leq\frac{d-1}{dk_{R}}\int_{\Omega}|\nabla f|^{2},

for all fC1(Ω)f\in C^{1}(\Omega) with Ωf𝑑S=0\int_{\partial\Omega}fdS=0, and which is of independent interest. The proof is based on a novel application of Reilly’s formula [28] which is also used for a complementary lower bound of σ\sigma independent of the interpolation approach stating that

σαmin(dk23d1α1α|Ω||Ω|,dd1kR),\sigma_{\alpha}\geq\min\left(\frac{dk_{2}}{3d-1}\frac{\alpha}{1-\alpha}\frac{|\partial\Omega|}{|\Omega|},\frac{d}{d-1}k_{R}\right),

but which is generally weaker for small values of α\alpha, c.f. Section 3.2.

The interpolation approach also yields a sufficient condition for the continuity of σα\sigma_{\alpha} at α{0,1}\alpha\in\{0,1\}, which in general may fail. In Section 2.2 we present sufficient conditions for continuity and discontinuity of σα\sigma_{\alpha} at {0,1}\{0,1\} which hints towards a phase transition in the associated family of variational problems.

We conclude with the discussion of two applications of the method in non-standard or singular situations, c.f. Sections 3.3 and 3.4.

2 An interpolation approach

2.1 Generalized framework

It will be convenient to work with a slight generalisation of the setup above. To this aim let Ω\Omega be an open domain in d\mathbb{R}^{d} or a Riemannian manifold with a piecewise smooth boundary Ω\partial\Omega. Let Σ\Sigma be a smooth compact and connected subset of Ω\partial\Omega. We denote by Σ\partial\Sigma the boundary of Σ\Sigma in the space Ω\partial\Omega, i.e. Σ=ΣΩ\Σ¯\partial\Sigma=\Sigma\cap\overline{\partial\Omega\backslash\Sigma}. We consider two probability measures λΩ\lambda_{\Omega} and λΣ\lambda_{\Sigma} with support Ω\Omega and Σ\Sigma, which are absolutely continuous with respect to the Lebesgue and the Hausdorff measure on Ω\Omega and Σ\Sigma, respectively.

Let 𝐷:C1(Ω)Γ0(Ω)\mathop{D}:C^{1}(\Omega)\mapsto\Gamma^{0}(\Omega) and 𝐷τ:C1(Ω)Γ0(Ω){\mathop{D}}^{\tau}:C^{1}(\partial\Omega)\mapsto\Gamma^{0}(\partial\Omega) denote given first order gradient operators mapping differentiable functions into (tangential) vector fields on Ω\Omega and on Ω\partial\Omega, respectively, and for α[0,1]\alpha\in[0,1] let

λα\displaystyle\lambda_{\alpha} :=αλΩ+(1α)λΣ,\displaystyle:=\alpha\lambda_{\Omega}+(1-\alpha)\lambda_{\Sigma},
α(f)\displaystyle\mathcal{E}_{\alpha}(f) :=αΩ𝐷f2𝑑λΩ+(1α)Σ𝐷τf2𝑑λΣ,f𝒟0,\displaystyle:=\alpha\int_{\Omega}\|\mathop{D}f\|^{2}d\lambda_{\Omega}+(1-\alpha)\int_{\Sigma}\|{\mathop{D}}^{\tau}f\|^{2}d\lambda_{\Sigma},\quad f\in\mathcal{D}_{0},

where 𝒟0𝒞1(Ω¯)\mathcal{D}_{0}\subset\mathcal{C}^{1}(\overline{\Omega}) is dense in C0(Ω)C_{0}(\Omega). We assume that for α[0,1]\alpha\in[0,1] the quadratic form (α,𝒟0)(\mathcal{E}_{\alpha},\mathcal{D}_{0}) is a pre-Dirichlet form on L2(Ω¯,λα)L^{2}(\overline{\Omega},\lambda_{\alpha}) whose closure we shall denote by (α,𝒟)(\mathcal{E}_{\alpha},\mathcal{D}), c.f. [15] for details. We wish to estimate from above σα1=Cα\sigma_{\alpha}^{-1}=C_{\alpha}, where CαC_{\alpha} is the optimal Poincaré constant given by

Cα:=supf𝒟0α(f)>0Varλαfα(f).C_{\alpha}:=\sup_{\begin{subarray}{c}f\in\mathcal{D}_{0}\\ \mathcal{E}_{\alpha}(f)>0\end{subarray}}\frac{\operatorname{Var}_{\lambda_{\alpha}}f}{\mathcal{E}_{\alpha}(f)}. (2.1)

In the interpolation method presented below it is assumed that CαC_{\alpha} are known or can be estimated at the two extremals α{0,1}\alpha\in\{0,1\}. For instance, when 𝐷=\mathop{D}=\nabla, 𝐷τ=τ{\mathop{D}}^{\tau}=\nabla^{\tau} are the standard gradient resp. tangential gradient operators and λΩ\lambda_{\Omega} and λΣ\lambda_{\Sigma} are normalized Lebesgue resp. Hausdorff measures on Ω\Omega and ΣΩ\Sigma\subset\partial\Omega, CΩ:=C1C_{\Omega}:=C_{1} is the optimal Poincaré constant associated to the Laplace operator on Ω\Omega with Neumann boundary conditions, whereas CΣ:=C0C_{\Sigma}:=C_{0} is the optimal Poincaré constant associated to the Laplace-Beltrami operator on Σ\Sigma with Neumann boundary conditions on Σ\partial\Sigma.

The following proposition establishes an estimate of CαC_{\alpha} in terms of CΩC_{\Omega} and CΣC_{\Sigma}.

Proposition 2.1.

Assume there exists constants KΣ,ΩK_{\Sigma,\Omega}, K1,K2K_{1},K_{2} such that for any f𝒟0f\in\mathcal{D}_{0}

VarλΣfKΣ,ΩΩ𝐷f2𝑑λΩ,\displaystyle\operatorname{Var}_{\lambda_{\Sigma}}f\leq K_{\Sigma,\Omega}\int_{\Omega}\|\mathop{D}f\|^{2}d\lambda_{\Omega}, (2.2)

and

(Ωf𝑑λΩΣf𝑑λΣ)2K1Ω𝐷f2𝑑λΩ+K2Σ𝐷τf2𝑑λΣ,\displaystyle\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}\leq K_{1}\int_{\Omega}\|\mathop{D}f\|^{2}d\lambda_{\Omega}+K_{2}\int_{\Sigma}\|{\mathop{D}}^{\tau}f\|^{2}d\lambda_{\Sigma}, (2.3)

then it holds for any α(0,1)\alpha\in(0,1),

Cαmax(CΩ+(1α)K1,αK2,(1α)KΣ,ΩCΣ+αCΩCΣ+α(1α)(KΣ,ΩK2+CΣK1)(1α)KΣ,Ω+αCΣ).\displaystyle C_{\alpha}\leq\max\left(C_{\Omega}+(1-\alpha)K_{1},\alpha K_{2},\frac{(1-\alpha)K_{\Sigma,\Omega}C_{\Sigma}+\alpha C_{\Omega}C_{\Sigma}+\alpha(1-\alpha)(K_{\Sigma,\Omega}K_{2}+C_{\Sigma}K_{1})}{(1-\alpha)K_{\Sigma,\Omega}+\alpha C_{\Sigma}}\right). (2.4)
Proof.

By definition of CΣC_{\Sigma} and by (2.2), for any f𝒟0f\in\mathcal{D}_{0}

VarλΣftKΣ,ΩΩ𝐷f2𝑑λΩ+(1t)CΣΣ𝐷τf2𝑑λΣ,\displaystyle\operatorname{Var}_{\lambda_{\Sigma}}f\leq tK_{\Sigma,\Omega}\int_{\Omega}\|\mathop{D}f\|^{2}d\lambda_{\Omega}+(1-t)C_{\Sigma}\int_{\Sigma}\|{\mathop{D}}^{\tau}f\|^{2}d\lambda_{\Sigma},

for any t[0,1]t\in[0,1]. Let α(0,1)\alpha\in(0,1). For any f𝒟0f\in\mathcal{D}_{0} and any t[0,1]t\in[0,1]

Varλαf\displaystyle\operatorname{Var}_{\lambda_{\alpha}}f =αVarλΩf+(1α)VarλΣf+α(1α)(Ωf𝑑λΩΣf𝑑λΣ)2\displaystyle=\alpha\operatorname{Var}_{\lambda_{\Omega}}f+(1-\alpha)\operatorname{Var}_{\lambda_{\Sigma}}f+\alpha(1-\alpha)\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}
(CΩ+(1α)tαKΣ,Ω+(1α)K1)αΩ𝐷f2𝑑λΩ\displaystyle\leq\left(C_{\Omega}+\frac{(1-\alpha)t}{\alpha}K_{\Sigma,\Omega}+(1-\alpha)K_{1}\right)\alpha\int_{\Omega}\|\mathop{D}f\|^{2}d\lambda_{\Omega}
+((1t)CΣ+αK2)(1α)Σ𝐷τf2𝑑λΣ.\displaystyle\quad+\left((1-t)C_{\Sigma}+\alpha K_{2}\right)(1-\alpha)\int_{\Sigma}\|{\mathop{D}}^{\tau}f\|^{2}d\lambda_{\Sigma}.

Therefore,

Cαinft[0,1]max(CΩ+(1α)tαKΣ,Ω+(1α)K1,(1t)CΣ+αK2).\displaystyle C_{\alpha}\leq\inf_{t\in[0,1]}\max\left(C_{\Omega}+\frac{(1-\alpha)t}{\alpha}K_{\Sigma,\Omega}+(1-\alpha)K_{1},(1-t)C_{\Sigma}+\alpha K_{2}\right).

For any positive constants a,b,c,da,b,c,d, we have

inft[0,1]max(a+bt,cdt)={aif ca<0,cdif ca>b+d,bc+adb+dif 0cab+d.\displaystyle\inf_{t\in[0,1]}\max\left(a+bt,c-dt\right)=\begin{cases}a&\text{if }c-a<0,\\ c-d&\text{if }c-a>b+d,\\ \frac{bc+ad}{b+d}&\text{if }0\leq c-a\leq b+d.\end{cases}

Therefore

Cα{CΩ+(1α)K1 if αK2(1α)K1+CΣCΩ<0,αK2 if αK2(1α)K1CΩ>1ααKΣ,Ω,(1α)KΣ,ΩCΣ+αCΩCΣ+α(1α)(KΣ,ΩK2+CΣK1)(1α)KΣ,Ω+αCΣif 0αK2(1α)K1+CΣCΩCΣ+1ααKΣ,Ω.\displaystyle C_{\alpha}\leq\begin{cases}C_{\Omega}+(1-\alpha)K_{1}&\text{\ \ if }\alpha K_{2}-(1-\alpha)K_{1}+C_{\Sigma}-C_{\Omega}<0,\\ \alpha K_{2}&\text{\ \ if }\alpha K_{2}-(1-\alpha)K_{1}-C_{\Omega}>\frac{1-\alpha}{\alpha}K_{\Sigma,\Omega},\\ \frac{(1-\alpha)K_{\Sigma,\Omega}C_{\Sigma}+\alpha C_{\Omega}C_{\Sigma}+\alpha(1-\alpha)(K_{\Sigma,\Omega}K_{2}+C_{\Sigma}K_{1})}{(1-\alpha)K_{\Sigma,\Omega}+\alpha C_{\Sigma}}&\begin{array}[]{r}\text{if }0\leq\alpha K_{2}-(1-\alpha)K_{1}+C_{\Sigma}-C_{\Omega}\\ \leq C_{\Sigma}+\frac{1-\alpha}{\alpha}K_{\Sigma,\Omega}.\end{array}\end{cases}

The last term is equivalent to the announced result. ∎

2.2 Continuity of CαC_{\alpha}

In general, the function αCα\alpha\mapsto C_{\alpha} might have discontinuities at α{0,1}\alpha\in\{0,1\} in which cases an upper bound for CαC_{\alpha} which interpolates continuously between C0C_{0} and C1C_{1} cannot exist. For example, when Ω=(0,b)×(0,1)2\Omega=(0,b)\times(0,1)\subset\mathbb{R}^{2} and Σ=[0,b]×{0}\Sigma=[0,b]\times\{0\}, straightforward computations yield

limα0Cα=max{CΣ,4π2},\lim_{\alpha\to 0}C_{\alpha}=\max\left\{C_{\Sigma},\frac{4}{\pi^{2}}\right\},

where CΣ=b2π2C_{\Sigma}=\frac{b^{2}}{\pi^{2}}. Hence αCα\alpha\mapsto C_{\alpha} is discontinuous at α=0\alpha=0 if and only if b<2b<2. – To generalize this to the framework of Section 2.1 let 𝒞01(Ω¯)={f𝒞1(Ω¯):f=0 on Σ}\mathcal{C}^{1}_{0}(\overline{\Omega})=\{f\in\mathcal{C}^{1}(\overline{\Omega}):f=0\mbox{ on }\Sigma\} and

C~0:=supf𝒞01(Ω¯)f non constantΩf2𝑑λΩΩ𝐷f2𝑑λΩ.\tilde{C}_{0}:=\sup_{\begin{subarray}{c}f\in\mathcal{C}^{1}_{0}(\overline{\Omega})\\ f\text{ non constant}\end{subarray}}\frac{\int_{\Omega}f^{2}d\lambda_{\Omega}}{\int_{\Omega}\|\mathop{D}f\|^{2}d\lambda_{\Omega}}.

(If 𝐷=\mathop{D}=\nabla, C~0\tilde{C}_{0} is the inverse of the spectral gap for Brownian motion on Ω\Omega with killing on Σ\Sigma and normal reflection at ΩΣ\partial\Omega\setminus\Sigma. ) We can then record the following statement as a partial corollary to Proposition 2.1.

Proposition 2.2.

In the setting of proposition 2.1 it holds that

lim¯α0CαC~0.\varliminf_{\alpha\to 0}C_{\alpha}\geq\tilde{C}_{0}.

In particular, if CΣ<C~0C_{\Sigma}<\tilde{C}_{0}, then αCα\alpha\mapsto C_{\alpha} is discontinuous at α=0\alpha=0. Conversely, if CΣCΩ+K1C_{\Sigma}\geq C_{\Omega}+K_{1} then αCα\alpha\mapsto C_{\alpha} is continuous at 0. If CΩK2C_{\Omega}\geq K_{2} continuity at 1 holds.

Proof.

To prove the second statement, take a non constant function g𝒞01(Ω¯)g\in\mathcal{C}^{1}_{0}(\overline{\Omega}) and estimate

lim¯α0Cα\displaystyle\varliminf_{\alpha\to 0}C_{\alpha} =lim¯α0supf𝒞1(Ω¯)f non constantVarλαfα(f)lim¯α0Varλαgα(g)\displaystyle=\varliminf_{\alpha\to 0}\sup_{\begin{subarray}{c}f\in\mathcal{C}^{1}(\overline{\Omega})\\ f\text{ non constant}\end{subarray}}\frac{\operatorname{Var}_{\lambda_{\alpha}}f}{\mathcal{E}_{\alpha}(f)}\geq\varliminf_{\alpha\to 0}\frac{\operatorname{Var}_{\lambda_{\alpha}}g}{\mathcal{E}_{\alpha}(g)}
=lim¯α0αVarλΩg+(1α)VarλΣg+α(1α)(Ωg𝑑λΩΣg𝑑λΣ)2αΩ𝐷g2𝑑λΩ+(1α)Σ𝐷τg2𝑑λΣ.\displaystyle=\varliminf_{\alpha\to 0}\frac{\alpha\operatorname{Var}_{\lambda_{\Omega}}g+(1-\alpha)\operatorname{Var}_{\lambda_{\Sigma}}g+\alpha(1-\alpha)\left(\int_{\Omega}gd\lambda_{\Omega}-\int_{\Sigma}gd\lambda_{\Sigma}\right)^{2}}{\alpha\int_{\Omega}\|\mathop{D}g\|^{2}d\lambda_{\Omega}+(1-\alpha)\int_{\Sigma}\|{\mathop{D}}^{\tau}g\|^{2}d\lambda_{\Sigma}}.

Since g=0g=0 on Σ\Sigma, we obtain

lim¯α0Cα\displaystyle\varliminf_{\alpha\to 0}C_{\alpha} lim¯α0αVarλΩg+α(1α)(Ωg𝑑λΩ)2αΩ𝐷g2𝑑λΩ=Ωg2𝑑λΩΩ𝐷g2𝑑λΩ.\displaystyle\geq\varliminf_{\alpha\to 0}\frac{\alpha\operatorname{Var}_{\lambda_{\Omega}}g+\alpha(1-\alpha)\left(\int_{\Omega}gd\lambda_{\Omega}\right)^{2}}{\alpha\int_{\Omega}\|\mathop{D}g\|^{2}d\lambda_{\Omega}}=\frac{\int_{\Omega}g^{2}d\lambda_{\Omega}}{\int_{\Omega}\|\mathop{D}g\|^{2}d\lambda_{\Omega}}.

Taking the supremum over g𝒞01(Ω¯)g\in\mathcal{C}^{1}_{0}(\overline{\Omega}) yields the first statement.

To prove the second assertion note that αCα\alpha\mapsto C_{\alpha} is the pointwise supremum of a family of continuous functions and therefore lower semi continuous. Thus CΣ=C0lim¯α0CαC_{\Sigma}=C_{0}\leq\varliminf_{\alpha\to 0}C_{\alpha}. If CΣCΩ+K1C_{\Sigma}\geq C_{\Omega}+K_{1}, the r.h.s. of inequality (2.4) converges to CΣC_{\Sigma} as α\alpha goes to 0, which implies that lim¯α0CαCΣ\varlimsup_{\alpha\to 0}C_{\alpha}\leq C_{\Sigma}. Similarly, if CΩK2C_{\Omega}\geq K_{2}, the r.h.s. of (2.4) converges to CΩC_{\Omega} as α\alpha goes ∎

Remark 2.3.

For smooth enough boundary the constant K2K_{2} can always be taken equal to zero, hence by proposition 2.2 continuity at α=1\alpha=1 holds. An example where a phase transition appears at α=0\alpha=0 is given in section 3.3. In section 3.4 we present an example where CΩ<K2C_{\Omega}<K_{2} but continuity of at α=1\alpha=1 can be established via Mosco-convergence [23] of the associated Dirichlet forms, see also [24].

3 Examples

3.1 Brownian motion on balls with sticky boundary diffusion

As our first example let Ω:=B1\Omega:=B_{1} be the unit ball in d\mathbb{R}^{d}, Σ=Ω\Sigma=\partial\Omega and 𝐷=\mathop{D}=\nabla and 𝐷τ=βτ{\mathop{D}}^{\tau}=\sqrt{\beta}\,\nabla^{\tau} with 𝒟0=C1(Ω¯)\mathcal{D}_{0}=C^{1}(\overline{\Omega}).

Proposition 3.1.

In the case when Ω=B1d\Omega=B_{1}\subset\mathbb{R}^{d} the optimal Poincaré constant of the generator (1.1) is bounded from above by

Cαmax(CΩ+(1α)d+14d2,4(1α)d+4αd2CΩ+α(1α)(d+1)4d(αd+(1α)β(d1))),\displaystyle C_{\alpha}\leq\max\left(C_{\Omega}+(1-\alpha)\frac{d+1}{4d^{2}},\frac{4(1-\alpha)d+4\alpha d^{2}C_{\Omega}+\alpha(1-\alpha)(d+1)}{4d(\alpha d+(1-\alpha)\beta(d-1))}\right), (3.1)

where α=γd+γ\alpha=\frac{\gamma}{d+\gamma} and CΩC_{\Omega} is the optimal Poincaré constant for reflecting Brownian motion on B1dB_{1}\subset\mathbb{R}^{d}.

Proof.

In order to apply Proposition 2.1, it is sufficient to compute the constants CΣC_{\Sigma}, KΣ,ΩK_{\Sigma,\Omega}, K1K_{1} and K2K_{2}. We claim that inequalities (2.2) and (2.3) holds with

CΣ=1β(d1),KΣ,Ω=1d,K1=d+14d2,K2=0.C_{\Sigma}=\frac{1}{\beta(d-1)},\quad K_{\Sigma,\Omega}=\frac{1}{d},\quad K_{1}=\frac{d+1}{4d^{2}},\quad K_{2}=0.

First, according to [31, Theorem 22.1], the first eigenvalue of the Laplace-Beltrami operator on the unit sphere of dimension d1d-1 is equal to d1d-1, thus CΣ=1β(d1)C_{\Sigma}=\frac{1}{\beta(d-1)}.

Moreover, according to [3, Theorem 4], for every f𝒞1(Ω)f\in\mathcal{C}^{1}(\partial\Omega) one has

(Ω|f|q𝑑λΣ)2qq2dΩu2𝑑λΩ+Ωf2𝑑λΣ,\left(\int_{\partial\Omega}|f|^{q}d\lambda_{\Sigma}\right)^{\frac{2}{q}}\leq\frac{q-2}{d}\int_{\Omega}\|\nabla u\|^{2}d\lambda_{\Omega}+\int_{\partial\Omega}f^{2}d\lambda_{\Sigma},

for 2q<2\leq q<\infty if d=2d=2 and 2q<2d2d22\leq q<\frac{2d-2}{d-2} if d3d\geq 3, where uu is the harmonic extension of ff to the unit ball Ω\Omega. It implies the logarithmic Sobolev inequality EntλΣ(f2)2dΩu2𝑑λΩ\operatorname{Ent}_{\lambda_{\Sigma}}(f^{2})\leq\frac{2}{d}\int_{\Omega}\|\nabla u\|^{2}d\lambda_{\Omega}. Repeating the proof of Proposition 5.1.3 in [2], we get VarλΣf1dΩu2𝑑λΩ\operatorname{Var}_{\lambda_{\Sigma}}f\leq\frac{1}{d}\int_{\Omega}\|\nabla u\|^{2}d\lambda_{\Omega}. Moreover, since the harmonic extension of ff is minimizing the energy functional 1\mathcal{E}_{1} under any function with boundary condition ff, the last inequality implies for any f𝒞1(Ω¯)f\in\mathcal{C}^{1}(\overline{\Omega})

VarλΣf1dΩf2𝑑λΩ,\displaystyle\operatorname{Var}_{\lambda_{\Sigma}}f\leq\frac{1}{d}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}, (3.2)

which implies KΣ,Ω=1dK_{\Sigma,\Omega}=\frac{1}{d}.

Furthermore, note that Ωf(y)λΣ(dy)=Ωf(πx)λΩ(dx)\int_{\partial\Omega}f(y)\lambda_{\Sigma}(dy)=\int_{\Omega}f(\pi_{x})\lambda_{\Omega}(dx), where πx=xx\pi_{x}=\frac{x}{\|x\|}, x0x\not=0. Hence, using Jensen’s inequality and polar coordinates

(Ωf𝑑λΩΩf𝑑λΣ)2\displaystyle\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\partial\Omega}fd\lambda_{\Sigma}\right)^{2} Ω(f(x)f(πx))2λΩ(dx)\displaystyle\leq\int_{\Omega}(f(x)-f(\pi_{x}))^{2}\lambda_{\Omega}(dx)
=1|Ω|Ω01(f(ry)f(y))2rd1𝑑r𝑑y\displaystyle=\frac{1}{|\Omega|}\int_{\partial\Omega}\int_{0}^{1}\left(f(ry)-f(y)\right)^{2}r^{d-1}drdy
=1|Ω|Ω01(r1ddsf(sy)𝑑s)2rd1𝑑r𝑑y\displaystyle=\frac{1}{|\Omega|}\int_{\partial\Omega}\int_{0}^{1}\left(\int_{r}^{1}\frac{d}{ds}f(sy)ds\right)^{2}r^{d-1}drdy
1|Ω|Ω01(1r)(r1(ddsf(sy))2𝑑s)rd1𝑑r𝑑y\displaystyle\leq\frac{1}{|\Omega|}\int_{\partial\Omega}\int_{0}^{1}(1-r)\left(\int_{r}^{1}\left(\frac{d}{ds}f(sy)\right)^{2}ds\right)r^{d-1}drdy
=1|Ω|Ω01[0s(1r)rd1𝑑r](ddsf(sy))2𝑑s𝑑y.\displaystyle=\frac{1}{|\Omega|}\int_{\partial\Omega}\int_{0}^{1}\left[\int_{0}^{s}(1-r)r^{d-1}dr\right]\left(\frac{d}{ds}f(sy)\right)^{2}dsdy.

We separately estimate

0s(1r)rd1𝑑r=(sds2d+1)sd1d+14d2sd1.\displaystyle\int_{0}^{s}(1-r)r^{d-1}dr=\left(\frac{s}{d}-\frac{s^{2}}{d+1}\right)s^{d-1}\leq\frac{d+1}{4d^{2}}s^{d-1}.

for any s[0,1]s\in[0,1]. Hence,

(Ωf𝑑λΩΩf𝑑λΣ)2\displaystyle\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\partial\Omega}fd\lambda_{\Sigma}\right)^{2} d+14d2|Ω|Ω01(f(sy)y)2sd1𝑑s\displaystyle\leq\frac{d+1}{4d^{2}|\Omega|}\int_{\partial\Omega}\int_{0}^{1}\left(\nabla f(sy)\cdot y\right)^{2}s^{d-1}ds
=d+14d2|Ω|Ω01f(sy)2sd1𝑑s𝑑y\displaystyle=\frac{d+1}{4d^{2}|\Omega|}\int_{\partial\Omega}\int_{0}^{1}\left\|\nabla f(sy)\right\|^{2}s^{d-1}dsdy
=d+14d2Ωf(x)2λΩ(dx).\displaystyle=\frac{d+1}{4d^{2}}\int_{\Omega}\|\nabla f(x)\|^{2}\lambda_{\Omega}(dx). (3.3)

which implies K1=d+14d2K_{1}=\frac{d+1}{4d^{2}} and K2=0K_{2}=0. ∎

For illustration, in d=2,d=2, we compare the bound from Proposition 3.1 for β=1,γ>0\beta=1,\gamma>0 to the optimal constant CαC_{\alpha} which will be computed numerically. To evaluate the bound (3.1), note that in this case

CΩ=1σΩ13.39,C_{\Omega}=\frac{1}{\sigma_{\Omega}}\approx\frac{1}{3.39}, (3.4)

where σΩ\sigma_{\Omega} is the smallest positive eigenvalue of the Laplace operator with Neumann boundary condition on the circle. It is given as the minimal positive solution to the equation Jm(γ)=0J_{m}^{\prime}(\sqrt{\gamma})=0, m0m\in\mathbb{N}_{0}, where JmJ_{m} is the Bessel function of the first kind of parameter mm, defined by Jm(x)=1π0πcos(mtxsint)𝑑tJ_{m}(x)=\frac{1}{\pi}\int_{0}^{\pi}\cos(mt-x\sin t)dt, x0x\geq 0. As a consequence, inequality (3.1) becomes

Cα8(1α)σΩ+16α+3α(1α)σΩ8(1+α)σΩ.\displaystyle C_{\alpha}\leq\frac{8(1-\alpha)\sigma_{\Omega}+16\alpha+3\alpha(1-\alpha)\sigma_{\Omega}}{8(1+\alpha)\sigma_{\Omega}}. (3.5)

For the numerical computation of CαC_{\alpha} one notes that the generator AαA_{\alpha} associated with α\mathcal{E}_{\alpha} is defined on D(Aα)𝒞2(Ω¯)D(A_{\alpha})\subset\mathcal{C}^{2}(\overline{\Omega}) as

Aαf=𝕀ΩΔf+𝕀Ω(Δτf2α1αfν),A_{\alpha}f=\mathbb{I}_{\Omega}\Delta f+\mathbb{I}_{\partial\Omega}\left(\Delta^{\tau}f-\frac{2\alpha}{1-\alpha}\frac{\partial f}{\partial\nu}\right),

where Δτ\Delta^{\tau} and ν\frac{\partial}{\partial\nu} denote the Laplace-Beltrami operator and the outer normal derivative on the circle Ω\partial\Omega. Hence, an eigenvector of Aα-A_{\alpha} for eigenvalue λ0\lambda\geq 0 is a function fD(Aα)f\in D(A_{\alpha}) such that

Aαf=λfinΩ.A_{\alpha}f=-\lambda f\quad\mbox{in}\ \ \Omega.

This equation is equivalent to the system of partial differential equations

{Δf=λfinΩ,Δτf2α1αfν=λfonΩ,\begin{cases}\Delta f=-\lambda f&\mbox{in}\ \ \Omega,\\ \Delta^{\tau}f-\frac{2\alpha}{1-\alpha}\frac{\partial f}{\partial\nu}=-\lambda f&\mbox{on}\ \ \partial\Omega,\end{cases}

which by the continuity of ff can be rewritten as

{Δf=λfinΩ,Δf=Δτf2α1αfνonΩ.\begin{cases}\Delta f=-\lambda f&\mbox{in}\ \ \Omega,\\ \Delta f=\Delta^{\tau}f-\frac{2\alpha}{1-\alpha}\frac{\partial f}{\partial\nu}&\mbox{on}\ \ \partial\Omega.\end{cases}

Passing to polar coordinates (x1,x2)=(rcosθ,rsinθ)Ω(x_{1},x_{2})=(r\cos\theta,r\sin\theta)\in\Omega in d=2d=2 and separating variables, we obtain the set of eigenfunctions {fm,lc,fm,ls}m,l0\{f_{m,l}^{c},f_{m,l}^{s}\}_{m,l\in\mathbb{N}_{0}},

fm,lc(x1,x2)=Jm(λm,lr)cos(mθ),m,l0,f_{m,l}^{c}(x_{1},x_{2})=J_{m}(\sqrt{\lambda_{m,l}}r)\cos(m\theta),\quad m,l\in\mathbb{N}_{0},
fm,ls(x1,x2)=Jm(λm,lr)sin(mθ),m,l0,f_{m,l}^{s}(x_{1},x_{2})=J_{m}(\sqrt{\lambda_{m,l}}r)\sin(m\theta),\quad m\in\mathbb{N},\ \ l\in\mathbb{N}_{0},

where λm,l\lambda_{m,l}, l0l\in\mathbb{N}_{0}, are countable family of positive solutions to the equation

λJm′′(λ)+1+α1αJm(λ)=0\sqrt{\lambda}J_{m}^{\prime\prime}(\sqrt{\lambda})+\frac{1+\alpha}{1-\alpha}J_{m}^{\prime}(\sqrt{\lambda})=0 (3.6)

for every m0m\in\mathbb{N}_{0}. Since the family {fm,lc,m,l0}{fm,ls,m0,l0}\{f_{m,l}^{c},\ m,l\in\mathbb{N}_{0}\}\cup\{f_{m,l}^{s},\ m\in\mathbb{N}_{0},\ l\in\mathbb{N}_{0}\} is dense in L2(Ω,λα)L_{2}(\Omega,\lambda_{\alpha}) and the operator AαA_{\alpha} is symmetric, the standard argument implies

Cα=1λα,,C_{\alpha}=\frac{1}{\lambda_{\alpha,\star}}, (3.7)

where λα,=minm,l0λm,l\lambda_{\alpha,\star}=\min\limits_{m,l\in\mathbb{N}_{0}}\lambda_{m,l}. The resulting curves are plotted in Figure 1.

Refer to caption
Figure 1: The blue curve represents αCα\alpha\mapsto C_{\alpha} the optimal Poincaré constant when Ω\Omega is the unit ball of 2\mathbb{R}^{2} with full boundary diffusion. The red curve is the upper estimate given by (3.5).

3.2 Smooth manifold with boundary

Let Ω\Omega be a smooth compact Riemannian manifold of dimension dd with piecewise smooth boundary Ω\partial\Omega. We denote by Ric\operatorname{Ric} the Ricci curvature of Ω\Omega and by II\mathrm{II} the second fundamental form on the boundary Ω\partial\Omega. Assume in this section that:

Assumption (M):kr>0,k2>0,Ric|ΩkRidandII|Ωk2id.\displaystyle\text{Assumption (M)}:\quad\quad\quad\quad\quad\exists k_{r}>0,k_{2}>0,\quad\operatorname{Ric}|_{\Omega}\geq k_{R}\operatorname{id}\quad\text{and}\quad\mathrm{II}|_{\partial\Omega}\geq k_{2}\operatorname{id}.

As before we consider Σ=Ω\Sigma=\partial\Omega, 𝐷=\mathop{D}=\nabla and 𝐷τ=τ{\mathop{D}}^{\tau}=\nabla^{\tau} with 𝒟0=C1(Ω¯)\mathcal{D}_{0}=C^{1}(\overline{\Omega}).

Proposition 3.2.

Under assumption (M), it holds that

Cαmax(CΩ+(1α)(d1)dkR,CΣdkR2(1α)dkR+αdk2kRCΩ+α(1α)(d1)k22(1α)+αk2CΣ)=:M1.\displaystyle C_{\alpha}\leq\max\left(C_{\Omega}+\frac{(1-\alpha)(d-1)}{dk_{R}},\frac{C_{\Sigma}}{dk_{R}}\cdot\frac{2(1-\alpha)dk_{R}+\alpha dk_{2}k_{R}C_{\Omega}+\alpha(1-\alpha)(d-1)k_{2}}{2(1-\alpha)+\alpha k_{2}C_{\Sigma}}\right)=:M_{1}. (3.8)

This statement is obtained via Proposition 2.1 and the two statements below.

Proposition 3.3.

Under assumption (M), inequality (2.3) is satisfied with K2=0K_{2}=0 and

K1=d1dkR.\displaystyle K_{1}=\frac{d-1}{dk_{R}}.
Proof.

Our goal is to obtain an lower bound of

inffC1(Ω¯)Ωf2𝑑λΩ(Ωf𝑑λΩΣf𝑑λΣ)2,\displaystyle\inf_{\begin{subarray}{c}f\in C^{1}(\overline{\Omega})\end{subarray}}\frac{\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}}{\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}},

where we recall that Σ=Ω\Sigma=\partial\Omega. We note that

inffC1(Ω¯)Ωf2𝑑λΩ(Ωf𝑑λΩΣf𝑑λΣ)2=inffC1(Ω¯)Σf𝑑λΣ=0Ωf2𝑑λΩ(Ωf𝑑λΩ)2inffC1(Ω¯)Σf𝑑λΣ=0Ωf2𝑑λΩΩf2𝑑λΩ=:σ.\displaystyle\inf_{\begin{subarray}{c}f\in C^{1}(\overline{\Omega})\end{subarray}}\frac{\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}}{\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}}=\inf_{\begin{subarray}{c}f\in C^{1}(\overline{\Omega})\\ \int_{\Sigma}fd\lambda_{\Sigma}=0\end{subarray}}\frac{\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}}{\left(\int_{\Omega}fd\lambda_{\Omega}\right)^{2}}\geq\inf_{\begin{subarray}{c}f\in C^{1}(\overline{\Omega})\\ \int_{\Sigma}fd\lambda_{\Sigma}=0\end{subarray}}\frac{\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}}{\int_{\Omega}f^{2}d\lambda_{\Omega}}=:\sigma.

Let fC1(Ω¯)f\in C^{1}(\overline{\Omega}) be a minimizer for σ\sigma. Then Σf𝑑λΣ=0\int_{\Sigma}fd\lambda_{\Sigma}=0 and

ΩfξdλΩ=σΩfξ𝑑λΩ\displaystyle\int_{\Omega}\nabla f\cdot\nabla\xi d\lambda_{\Omega}=\sigma\int_{\Omega}f\xi d\lambda_{\Omega}

for each ξC1(Ω¯)\xi\in C^{1}(\overline{\Omega}) with Σξ𝑑λΣ=0\int_{\Sigma}\xi d\lambda_{\Sigma}=0. By integration by parts, the latter equality is equivalent to

ΩΔfξ𝑑λΩ+|Σ||Ω|Σfνξ𝑑λΣ=σΩfξ𝑑λΩ\displaystyle-\int_{\Omega}\Delta f\xi d\lambda_{\Omega}+\frac{|\Sigma|}{|\Omega|}\int_{\Sigma}\frac{\partial f}{\partial\nu}\xi d\lambda_{\Sigma}=\sigma\int_{\Omega}f\xi d\lambda_{\Omega}

for each ξC1(Ω¯)\xi\in C^{1}(\overline{\Omega}) satisfying Σξ𝑑λΣ=0\int_{\Sigma}\xi d\lambda_{\Sigma}=0. In particular, choosing ξC0(Ω)\xi\in C^{\infty}_{0}(\Omega) (which obviously satisfies Σξ𝑑λΣ=0\int_{\Sigma}\xi d\lambda_{\Sigma}=0), we get that ff should satisfy Δf=σf-\Delta f=\sigma f in Ω\Omega. Hence Σfνξ𝑑λΣ=0\int_{\Sigma}\frac{\partial f}{\partial\nu}\xi d\lambda_{\Sigma}=0 for each ξ\xi with zero mean, so it follows that Σfν(ξΣξ𝑑λΣ)𝑑λΣ=0\int_{\Sigma}\frac{\partial f}{\partial\nu}\left(\xi-\int_{\Sigma}\xi d\lambda_{\Sigma}\right)d\lambda_{\Sigma}=0 for every ξC1(Ω¯)\xi\in C^{1}(\overline{\Omega}), which is equivalent to

Σ(fνΣfν𝑑λΣ)ξ𝑑λΣ=0\displaystyle\int_{\Sigma}\left(\frac{\partial f}{\partial\nu}-\int_{\Sigma}\frac{\partial f}{\partial\nu}d\lambda_{\Sigma}\right)\xi d\lambda_{\Sigma}=0

for every ξC1(Ω¯)\xi\in C^{1}(\overline{\Omega}). It follows that fν\frac{\partial f}{\partial\nu} is constant on Σ\Sigma. Therefore, ff satisfies

{Δf=σfinΩ,fνconΩ,Σf𝑑λΣ=0,\begin{cases}\Delta f=-\sigma f&\mbox{in}\ \ \Omega,\\ \frac{\partial f}{\partial\nu}\equiv c&\mbox{on}\ \ \partial\Omega,\\ \int_{\Sigma}fd\lambda_{\Sigma}=0,\end{cases} (3.9)

for some constant cc.

Moreover, recall Reilly’s formula (see [28])

Ω((Δf)22f2)𝑑x\displaystyle\int_{\Omega}\left((\Delta f)^{2}-\|\nabla^{2}f\|^{2}\right)dx =ΩRic(f,f)𝑑x\displaystyle=\int_{\Omega}\operatorname{Ric}(\nabla f,\nabla f)dx (3.10)
+Σ(H(fν)2+II(τf,τf)+2Δτffν)𝑑S\displaystyle\quad+\int_{\Sigma}\left(H(\frac{\partial f}{\partial\nu})^{2}+\mathrm{II}(\nabla^{\tau}f,\nabla^{\tau}f)+2\Delta^{\tau}f\frac{\partial f}{\partial\nu}\right)dS

where dxdx and dSdS denote the Riemannian volume resp. surface measure on Ω\Omega and Ω\partial\Omega, 2f\nabla^{2}f is the Hessian of ff and HH is the mean curvature of Σ\Sigma (i.e. the trace of II\mathrm{II}). Since ff satisfies (3.9),

Ω(Δf)2𝑑x=σΩfΔf𝑑x\displaystyle\int_{\Omega}(\Delta f)^{2}dx=-\sigma\int_{\Omega}f\Delta fdx =σΩf2𝑑xσΣfνf𝑑S\displaystyle=\sigma\int_{\Omega}\|\nabla f\|^{2}dx-\sigma\int_{\Sigma}\frac{\partial f}{\partial\nu}fdS
=σΩf2𝑑xσcΣf𝑑S=σΩf2𝑑x,\displaystyle=\sigma\int_{\Omega}\|\nabla f\|^{2}dx-\sigma c\int_{\Sigma}fdS=\sigma\int_{\Omega}\|\nabla f\|^{2}dx,

because Σf𝑑S=|Σ|Σf𝑑λΣ=0\int_{\Sigma}fdS=|\Sigma|\int_{\Sigma}fd\lambda_{\Sigma}=0. Furthermore, note that 2f2=i,j(ij2f)2i=1d(ii2f)21d(i=1dii2f)2=1d(Δf)2\|\nabla^{2}f\|^{2}=\sum_{i,j}(\partial_{ij}^{2}f)^{2}\geq\sum_{i=1}^{d}(\partial_{ii}^{2}f)^{2}\geq\frac{1}{d}(\sum_{i=1}^{d}\partial_{ii}^{2}f)^{2}=\frac{1}{d}(\Delta f)^{2}. Therefore, the l.h.s. of (3.10) is bounded by

Ω((Δf)22f2)𝑑xd1dΩ(Δf)2𝑑xd1dσΩf2𝑑x.\displaystyle\int_{\Omega}\left((\Delta f)^{2}-\|\nabla^{2}f\|^{2}\right)dx\leq\frac{d-1}{d}\int_{\Omega}(\Delta f)^{2}dx\leq\frac{d-1}{d}\sigma\int_{\Omega}\|\nabla f\|^{2}dx.

On the other hand, by assumption (M), H0H\geq 0, II(τf,τf)0\mathrm{II}(\nabla^{\tau}f,\nabla^{\tau}f)\geq 0 and

ΩRic(f,f)𝑑xkRΩf2𝑑x.\displaystyle\int_{\Omega}\operatorname{Ric}(\nabla f,\nabla f)dx\geq k_{R}\int_{\Omega}\|\nabla f\|^{2}dx.

Since

ΣΔτffν𝑑S=cΣΔτf𝑑S=0\displaystyle\int_{\Sigma}\Delta^{\tau}f\frac{\partial f}{\partial\nu}dS=c\int_{\Sigma}\Delta^{\tau}fdS=0

the r.h.s. of (3.10) is bounded from below by kRΩf2𝑑xk_{R}\int_{\Omega}\|\nabla f\|^{2}dx. It turns out that

d1dσΩf2𝑑xkRΩf2𝑑x,\displaystyle\frac{d-1}{d}\sigma\int_{\Omega}\|\nabla f\|^{2}dx\geq k_{R}\int_{\Omega}\|\nabla f\|^{2}dx,

which implies that σdd1kR\sigma\geq\frac{d}{d-1}k_{R}. It follows that inequality (2.3) holds with K1=d1dkRK_{1}=\frac{d-1}{dk_{R}}. ∎

Remark 3.4.

Instead of using K1K_{1} from Proposition 3.3 another admissible choice is

K1=|Ω||Ω|B2(1+CΩ)<,K_{1}^{\prime}=\frac{|\Omega|}{|\partial\Omega|}B^{2}(1+C_{\Omega})<\infty,

where BB is the optimal Sobolev trace constant of Ω\Omega, i.e. the norm of the embedding H1,2(Ω)L2(Ω)H^{1,2}(\Omega)\hookrightarrow L^{2}(\partial\Omega). B2B^{-2} is the first nontrivial eigenvalue of a Steklov-type eigenvalue problem

{Δf+f=0 in Ωfν=σf on Ω,\begin{cases}-\Delta f+f=0&\mbox{ in }\Omega\\ \frac{\partial f}{\partial\nu}=\sigma f&\mbox{ on }\partial\Omega,\end{cases}

for which however explicit lower bounds in terms of the geometry of Ω\Omega seem yet unknown [4, 5, 11, 21, 29].

Proposition 3.5.

Under assumption (M), inequality (2.2) holds with KΣ,Ω=2k2K_{\Sigma,\Omega}=\frac{2}{k_{2}}.

Proof.

The optimal choice for KΣ,ΩK_{\Sigma,\Omega} is σ1\sigma^{-1}, where σ\sigma given by

σ=inffC1(Ω¯)Σf𝑑λΣ=0Ωf2𝑑λΩ(Σf2𝑑λΣ)2\sigma=\inf_{\begin{subarray}{c}f\in C^{1}(\overline{\Omega})\\ \int_{\Sigma}fd\lambda_{\Sigma}=0\end{subarray}}\frac{\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}}{\left(\int_{\Sigma}f^{2}d\lambda_{\Sigma}\right)^{2}}

is the first nontrivial eigenvalue of the Steklov-problem c.f. [12]

{Δf=0 in Ω,fν=σf on Ω.\begin{cases}\Delta f=0&\mbox{ in }\Omega,\\ \frac{\partial f}{\partial\nu}=\sigma f&\mbox{ on }\partial\Omega.\end{cases}

Escobar [9] showed σk22\sigma\geq\frac{k_{2}}{2} in this case. ∎

Alternatively, we obtain another upper bound for CαC_{\alpha} by a direct application of Reilly’s formula.

Proposition 3.6.

Under assumption (M) it holds that

Cαmax((3d1)(1α)dαk2|Ω||Ω|,d1dkR)=:M2.\displaystyle C_{\alpha}\leq\max\left(\frac{(3d-1)(1-\alpha)}{d\alpha k_{2}}\frac{|\Omega|}{|\partial\Omega|},\frac{d-1}{dk_{R}}\right)=:M_{2}. (3.11)
Proof.

We estimate equivalently from below the first nontrivial eigenvalue σ=Cα1\sigma=C_{\alpha}^{-1} for the problem

{Δf+σf=0 in ΩΔτfγfν+σf=0 on Ω,\begin{cases}\Delta f+\sigma f=0&\mbox{ in }\Omega\\ \Delta^{\tau}f-\gamma\frac{\partial f}{\partial\nu}+\sigma f=0&\mbox{ on }\partial\Omega,\end{cases}

where γ=α1α|Ω||Ω|\gamma=\frac{\alpha}{1-\alpha}\frac{|\partial\Omega|}{|\Omega|}. As in the proof of Proposition 3.3 we apply Reilly’s formula (3.10) to the corresponding eigenfunction ff. In this case, for the l.h.s. we estimae

Ω((Δf)22f2)𝑑x\displaystyle\int_{\Omega}\left((\Delta f)^{2}-\|\nabla^{2}f\|^{2}\right)dx d1dΩ(Δf)2𝑑x=d1dσΩfΔf𝑑x\displaystyle\leq\frac{d-1}{d}\int_{\Omega}(\Delta f)^{2}dx=-\frac{d-1}{d}\sigma\int_{\Omega}f\Delta fdx
=d1dσΩf2𝑑xd1dσΣfνf𝑑S\displaystyle=\frac{d-1}{d}\sigma\int_{\Omega}\|\nabla f\|^{2}dx-\frac{d-1}{d}\sigma\int_{\Sigma}\frac{\partial f}{\partial\nu}fdS
=d1dσΩf2𝑑xd1dσγΣ(Δτf+σf)f𝑑S\displaystyle=\frac{d-1}{d}\sigma\int_{\Omega}\|\nabla f\|^{2}dx-\frac{d-1}{d}\frac{\sigma}{\gamma}\int_{\Sigma}(\Delta^{\tau}f+{\sigma}f)fdS
=d1dσΩf2𝑑x+d1dσγΣτf2𝑑Sd1dσ2γΣf2𝑑S\displaystyle=\frac{d-1}{d}\sigma\int_{\Omega}\|\nabla f\|^{2}dx+\frac{d-1}{d}\frac{\sigma}{\gamma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS-\frac{d-1}{d}\frac{\sigma^{2}}{\gamma}\int_{\Sigma}f^{2}dS
d1dσΩf2𝑑x+d1dσγΣτf2𝑑S.\displaystyle\leq\frac{d-1}{d}\sigma\int_{\Omega}\|\nabla f\|^{2}dx+\frac{d-1}{d}\frac{\sigma}{\gamma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS.

Since

ΣfνΔτf𝑑S\displaystyle\int_{\Sigma}\frac{\partial f}{\partial\nu}\Delta^{\tau}fdS =1γΣ(Δτf+σf)Δτf𝑑S\displaystyle=\frac{1}{\gamma}\int_{\Sigma}(\Delta^{\tau}f+\sigma f)\Delta^{\tau}fdS
=1γΣ(Δτf)2𝑑SσγΣτf2𝑑SσγΣτf2𝑑S\displaystyle=\frac{1}{\gamma}\int_{\Sigma}(\Delta^{\tau}f)^{2}dS-\frac{\sigma}{\gamma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS\geq-\frac{\sigma}{\gamma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS

the r.h.s. of (3.10) is bounded from below by

kRΩf2𝑑x\displaystyle k_{R}\int_{\Omega}\|\nabla f\|^{2}dx 2σγΣτf2𝑑S+Σh|fν|2𝑑S+k2Στf2𝑑S\displaystyle-\frac{2\sigma}{\gamma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS+\int_{\Sigma}h|\frac{\partial f}{\partial\nu}|^{2}dS+k_{2}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS
kRΩf2𝑑x2σγΣτf2𝑑S+k2Στf2𝑑S.\displaystyle\geq k_{R}\int_{\Omega}\|\nabla f\|^{2}dx-\frac{2\sigma}{\gamma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS+k_{2}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS.

Combining the two bounds for (3.10) yields

(d1dσkR)Ωf2𝑑x(k23d1dσγ)Στf2𝑑S,\left(\frac{d-1}{d}\sigma-k_{R}\right)\int_{\Omega}\|\nabla f\|^{2}dx\geq\left(k_{2}-\frac{3d-1}{d}\frac{\sigma}{\gamma}\right)\int_{\Sigma}\|\nabla^{\tau}f\|^{2}dS,

which implies that either

k23d1dσγ0, i.e. σdk2γ3d1k_{2}-\frac{3d-1}{d}\frac{\sigma}{\gamma}\leq 0,\quad\mbox{ i.e. }\quad\sigma\geq\frac{dk_{2}\gamma}{3d-1}

or

d1dσkR0, i.e. σdd1kr.\frac{d-1}{d}\sigma-k_{R}\geq 0,\quad\mbox{ i.e. }\quad\sigma\geq\frac{d}{d-1}k_{r}.

Consequently,

σmin(dk2γ3d1,dd1kR).\sigma\geq\min\left(\frac{dk_{2}\gamma}{3d-1},\frac{d}{d-1}k_{R}\right).

Corollary 3.7.

Under assumption (M), it holds that

Cαmin(M1,M2),\displaystyle C_{\alpha}\leq\min(M_{1},M_{2}),

where M1=M1(α)M_{1}=M_{1}(\alpha) and M2=M2(α)M_{2}=M_{2}(\alpha) are defined by (3.8) and (3.11), respectively.

When α\alpha goes to 0, M1M_{1} tends to max(CΩ,d1dkR,CΣ)\max(C_{\Omega},\frac{d-1}{dk_{R}},C_{\Sigma}) and M2M_{2} tends to ++\infty, so the estimation via the interpolation method is always stronger. When α\alpha goes to 11, M1M_{1} tends to CΩC_{\Omega} and M2M_{2} tends to d1dkR\frac{d-1}{dk_{R}}, so the relative strength of each method depends on the values of CΩC_{\Omega}, dd and kRk_{R}.

3.3 Brownian motion on balls with partial sticky reflecting boundary diffusion

As in Section 3.1, let Ω:=B1\Omega:=B_{1} be the unit ball of 2\mathbb{R}^{2}. Now, define for a fixed δ(0,1)\delta\in(0,1)

Σ={(cosθ,sinθ)Ω:δπθδπ},ΣN:=Ω\Σ.\Sigma=\{(\cos\theta,\sin\theta)\in\partial\Omega:-\delta\pi\leq\theta\leq\delta\pi\},\quad\quad\Sigma_{\operatorname{N}}:=\partial\Omega\backslash\Sigma.
Proposition 3.8.

It holds that

Cαmax(CΩ+(1α)K1(δ),4(1α)δ2+8αδ3CΩ+8α(1α)δ3K1(δ)(1α)+8αδ3),\displaystyle C_{\alpha}\leq\max\left(C_{\Omega}+(1-\alpha)K_{1}(\delta),\frac{4(1-\alpha)\delta^{2}+8\alpha\delta^{3}C_{\Omega}+8\alpha(1-\alpha)\delta^{3}K_{1}(\delta)}{(1-\alpha)+8\alpha\delta^{3}}\right), (3.12)

where CΩ=1σΩ13.39C_{\Omega}=\frac{1}{\sigma_{\Omega}}\approx\frac{1}{3.39} and K1(δ)=(1δπ+143δ)2K_{1}(\delta)=\left(\sqrt{1-\delta}\pi+\frac{1}{4}\sqrt{\frac{3}{\delta}}\right)^{2}.

As previously, we will start by computing the needed constants CΩC_{\Omega}, CΣC_{\Sigma}, KΣ,ΩK_{\Sigma,\Omega}, K1K_{1} and K2K_{2}. The first constant, CΩ=1σΩ13.39C_{\Omega}=\frac{1}{\sigma_{\Omega}}\approx\frac{1}{3.39}, remains unchanged.

Lemma 3.9.

The following inequalities hold true

VarλΣf\displaystyle\operatorname{Var}_{\lambda_{\Sigma}}f CΣΣτf2𝑑λΣ,\displaystyle\leq C_{\Sigma}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}d\lambda_{\Sigma}, (3.13)
VarλΣf\displaystyle\operatorname{Var}_{\lambda_{\Sigma}}f KΣ,ΩΩf2𝑑λΩ,\displaystyle\leq K_{\Sigma,\Omega}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}, (3.14)

where CΣ=4δ2C_{\Sigma}=4\delta^{2} and KΣ,Ω=12δK_{\Sigma,\Omega}=\frac{1}{2\delta}.

Proof.

Inequality (3.13) corresponds to the Poincaré inequality of the Laplacian on the one-dimensional interval [δπ,δπ][-\delta\pi,\delta\pi] with Neumann boundary conditions. It is well known (see [2, Prop. 4.5.5]) that the optimal Poincaré constant is given by CΣ=4δ2C_{\Sigma}=4\delta^{2}.

Moreover, let us decompose the normalized Hausdorff measure λ\lambda_{\partial} on the sphere Ω\partial\Omega into the normalized Hausdorff measure λΣ\lambda_{\Sigma} on Σ\Sigma and the normalized Hausdorff measure λN\lambda_{\operatorname{N}} on ΣN\Sigma_{\operatorname{N}}: λ=δλΣ+(1δ)λN\lambda_{\partial}=\delta\lambda_{\Sigma}+(1-\delta)\lambda_{\operatorname{N}}. Therefore

Varλf=δVarλΣf+(1δ)VarλNf+δ(1δ)(Σf𝑑λΣΣNf𝑑λN)2δVarλΣf,\displaystyle\operatorname{Var}_{\lambda_{\partial}}f=\delta\operatorname{Var}_{\lambda_{\Sigma}}f+(1-\delta)\operatorname{Var}_{\lambda_{\operatorname{N}}}f+\delta(1-\delta)\left(\int_{\Sigma}fd\lambda_{\Sigma}-\int_{\Sigma_{\operatorname{N}}}fd\lambda_{\operatorname{N}}\right)^{2}\geq\delta\operatorname{Var}_{\lambda_{\Sigma}}f,

Furthermore, recall that by inequality (3.2), for any f𝒞1(Ω¯)f\in\mathcal{C}^{1}(\overline{\Omega}), Varλf12Ωf2𝑑λΩ\operatorname{Var}_{\lambda_{\partial}}f\leq\frac{1}{2}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}. It implies (3.14). ∎

Lemma 3.10.

It holds that

(Ωf𝑑λΩΣf𝑑λΣ)2K1(δ)Ωf2𝑑λΩ\displaystyle\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}\leq K_{1}(\delta)\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}

with K1(δ)=(1δπ+143δ)2K_{1}(\delta)=\left(\sqrt{1-\delta}\pi+\frac{1}{4}\sqrt{\frac{3}{\delta}}\right)^{2}.

Proof.

For every xΩ\{0}x\in\Omega\backslash\{0\} with polar coordinates (r,θ)(r,\theta), r(0,1)r\in(0,1), θ(π,π]\theta\in(-\pi,\pi], denote by pxp_{x} the point of coordinates (1,δθ)(1,\delta\theta) on Σ\Sigma. Obviously, Σf(y)λΣ(dy)=Ωf(px)λΩ(dx)\int_{\Sigma}f(y)\lambda_{\Sigma}(dy)=\int_{\Omega}f(p_{x})\lambda_{\Omega}(dx) and by Jensen’s inequality

I:=(Ωf𝑑λΩΣf𝑑λΣ)2Ω(f(x)f(px))2λΩ(dx).\displaystyle I:=\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}\leq\int_{\Omega}\left(f(x)-f(p_{x})\right)^{2}\lambda_{\Omega}(dx).

Define g(r,θ):=f(rcos(θ),rsin(θ))g(r,\theta):=f(r\cos(\theta),r\sin(\theta)). Then

I\displaystyle I 1π01ππ(g(r,θ)g(1,δθ))2r𝑑r𝑑θ(J1+J2)2,\displaystyle\leq\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}(g(r,\theta)-g(1,\delta\theta))^{2}rdrd\theta\leq(\sqrt{J_{1}}+\sqrt{J_{2}})^{2}, (3.15)

where J1=1π01ππ(g(r,θ)g(r,δθ))2r𝑑r𝑑θJ_{1}=\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}(g(r,\theta)-g(r,\delta\theta))^{2}rdrd\theta and J2=1π01ππ(g(r,δθ)g(1,δθ))2r𝑑r𝑑θJ_{2}=\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}(g(r,\delta\theta)-g(1,\delta\theta))^{2}rdrd\theta. On the one hand

J1\displaystyle J_{1} =1π01ππ(δθθgθ(r,u)𝑑u)2r𝑑r𝑑θ1δπ01ππ|θ|ππ(gθ)2(r,u)𝑑ur𝑑r𝑑θ\displaystyle=\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}\left(\int_{\delta\theta}^{\theta}\frac{\partial g}{\partial\theta}(r,u)du\right)^{2}rdrd\theta\leq\frac{1-\delta}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}|\theta|\int_{-\pi}^{\pi}\left(\frac{\partial g}{\partial\theta}\right)^{2}(r,u)du\;rdrd\theta
(1δ)π21π01ππ(1rgθ)2(r,u)𝑑ur𝑑r(1δ)π2Ωf2𝑑λΩ.\displaystyle\leq(1-\delta)\pi^{2}\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}\left(\frac{1}{r}\frac{\partial g}{\partial\theta}\right)^{2}(r,u)du\;rdr\leq(1-\delta)\pi^{2}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}. (3.16)

On the other hand

J2\displaystyle J_{2} 1π01ππ(1r)r1(gr)2(s,δθ)𝑑sr𝑑r𝑑θ1π01ππ(gr)2(s,δθ)0s(1r)r𝑑r𝑑s𝑑θ.\displaystyle\leq\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}(1-r)\int_{r}^{1}\left(\frac{\partial g}{\partial r}\right)^{2}(s,\delta\theta)ds\;rdrd\theta\leq\frac{1}{\pi}\int_{0}^{1}\int_{-\pi}^{\pi}\left(\frac{\partial g}{\partial r}\right)^{2}(s,\delta\theta)\int_{0}^{s}(1-r)rdrdsd\theta.

For every s[0,1]s\in[0,1], 0s(1r)r𝑑r=s22s333s16\int_{0}^{s}(1-r)rdr=\frac{s^{2}}{2}-\frac{s^{3}}{3}\leq\frac{3s}{16}, thus

J2\displaystyle J_{2} 316δπ01δπδπ(gr)2(s,u)s𝑑s𝑑u316δΩf2𝑑λΩ.\displaystyle\leq\frac{3}{16\delta\pi}\int_{0}^{1}\int_{-\delta\pi}^{\delta\pi}\left(\frac{\partial g}{\partial r}\right)^{2}(s,u)sdsdu\leq\frac{3}{16\delta}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}. (3.17)

The proof of the lemma is completed by putting together (3.15), (3.16) and (3.17). ∎

Proof of Proposition 3.8.

We apply Proposition 2.1 with CΩ=1σΩC_{\Omega}=\frac{1}{\sigma_{\Omega}}, CΣ=4δ2C_{\Sigma}=4\delta^{2}, KΣ,Ω=12δK_{\Sigma,\Omega}=\frac{1}{2\delta}, K1(δ)=(1δπ+143δ)2K_{1}(\delta)=\left(\sqrt{1-\delta}\pi+\frac{1}{4}\sqrt{\frac{3}{\delta}}\right)^{2} and K2=0K_{2}=0. ∎

Refer to caption
(a) δ=0.5\delta=0.5
Refer to caption
(b) δ=0.9\delta=0.9
Figure 2: The above two figures show the upper estimate given by the r.h.s of (3.12). In the case δ=0.9\delta=0.9 (Figure 2b), the curve interpolates between the extremal constants CΣC_{\Sigma} and CΩC_{\Omega}, as opposed to the half-sphere case (Figure 2a).

For δ\delta sufficiently large, the map αCα\alpha\mapsto C_{\alpha} is continuous at α=0\alpha=0. Indeed, by Proposition 2.2, a sufficient condition is CΣ(δ)>CΩ+K1(δ)C_{\Sigma}(\delta)>C_{\Omega}+K_{1}(\delta), that is

4δ2>1σΩ+(1δπ+143δ)2,4\delta^{2}>\frac{1}{\sigma_{\Omega}}+\left(\sqrt{1-\delta}\pi+\frac{1}{4}\sqrt{\frac{3}{\delta}}\right)^{2},

which is satisfied for any δ0.862\delta\geq 0.862.

3.4 Ball with a needle

Our final example is the unit ball Ω=B1\Omega=B_{1} of 2\mathbb{R}^{2} with a needle \mathcal{L} of length LL attached to one point of the boundary, i.e.  :={(x,0):1xL+1}\mathcal{L}:=\{(x,0):1\leq x\leq L+1\}, see Figure 3. The attachment point and the endpoint of the needle are denoted by x0:=(1,0)x_{0}:=(1,0) and xL=(L+1,0)x_{L}=(L+1,0), respectively.

Refer to caption
Figure 3: The ball (in green) is denoted by Ω\Omega, the boundary of the ball is denoted by Ω\partial\Omega and the needle (in blue) is denoted by \mathcal{L}.

In that setting, we define Ω¯=B1¯\overline{\Omega}=\overline{B_{1}}\cup\mathcal{L}, Σ=B1\Sigma=\partial B_{1}\cup\mathcal{L} and

λα=αλΩ+(1α)λΣ,\lambda_{\alpha}=\alpha\lambda_{\Omega}+(1-\alpha)\lambda_{\Sigma},

where λΩ\lambda_{\Omega} is as previously the normalized Lebesgue measure on Ω\Omega and λΣ=2π2π+Lλ+L2π+Lλ\lambda_{\Sigma}=\frac{2\pi}{2\pi+L}\lambda_{\partial}+\frac{L}{2\pi+L}\lambda_{\mathcal{L}}, with λ\lambda_{\partial} and λ\lambda_{\mathcal{L}} being the normalized Hausdorff measures on Ω\partial\Omega and \mathcal{L}, respectively. We choose

𝒟0={fC0(Ω¯))C1(Ω¯{x0})|fe1+fe2+fe3=0 at x0},\mathcal{D}_{0}=\left\{f\in C_{0}(\overline{\Omega}))\cap C^{1}(\overline{\Omega}\setminus\{x_{0}\})\,|\,\frac{\partial f}{\partial e_{1}}+\frac{\partial f}{\partial e_{2}}+\frac{\partial f}{\partial e_{3}}=0\mbox{ at }x_{0}\right\},

where e1=(0,1)e_{1}=(0,1), e2=(0,1)e_{2}=(0,-1) and e3=(1,0)e_{3}=(1,0) are the three ”tangent” vectors to Σ\Sigma at point x0x_{0}, and 𝐷:=\mathop{D}:=\nabla, 𝐷τ:=βτ{\mathop{D}}^{\tau}:=\sqrt{\beta}\nabla^{\tau}, which is well defined in Σ{x0}\Sigma\setminus\{x_{0}\}. With this choice, for α[0,1]\alpha\in[0,1] (α,𝒟0)(\mathcal{E}_{\alpha},\mathcal{D}_{0}) is a pre-Dirichlet form on L2(Ω¯,λα)L^{2}(\overline{\Omega},\lambda_{\alpha}), whose closure generates Brownian motion on Ω\Omega with sticky boundary diffusion on Σ\Sigma, i.e. whose generator is given by

Aα(f)=Δf𝕀Ω+βΔΣf𝕀Σα1α2π+Lπfν𝕀Ω,A_{\alpha}(f)=\Delta f\mathbb{I}_{\Omega}+\beta\Delta_{\Sigma}f\mathbb{I}_{\Sigma}-\frac{\alpha}{1-\alpha}\frac{2\pi+L}{\pi}\frac{\partial f}{\partial\nu}\mathbb{I}_{\partial\Omega},

with ΔΣ\Delta_{\Sigma} being the generator of the canonical diffusion on Σ\Sigma with reflecting boundary condition at xLx_{L}. As before, the optimal Poincaré constant CαC_{\alpha} for AαA_{\alpha} is given by

Cα:=supf𝒟0α(f)>0Varλαfα(f),\displaystyle C_{\alpha}:=\sup_{\begin{subarray}{c}f\in\mathcal{D}_{0}\\ \mathcal{E}_{\alpha}(f)>0\end{subarray}}\frac{\operatorname{Var}_{\lambda_{\alpha}}f}{\mathcal{E}_{\alpha}(f)},

and let CΩ:=C1C_{\Omega}:=C_{1} and CΣ:=C0C_{\Sigma}:=C_{0}. In this case the following estimate is obtained.

Proposition 3.11.
Cαmax(1σΩ+38(1α),1βγL+αL2(π+L)β(2π+L)),\displaystyle C_{\alpha}\leq\max\left(\frac{1}{\sigma_{\Omega}}+\frac{3}{8}(1-\alpha),\frac{1}{\beta\gamma_{L}}+\alpha\frac{L^{2}(\pi+L)}{\beta(2\pi+L)}\right),

where γL>0\gamma_{L}>0 is the smallest positive solution to

2cos(γL)(1cos(γ2π))+sin(γL)sin(γ2π)=0.\displaystyle 2\cos(\sqrt{\gamma}L)(1-\cos(\sqrt{\gamma}2\pi))+\sin(\sqrt{\gamma}L)\sin(\sqrt{\gamma}2\pi)=0. (3.18)

Note that γL1\gamma_{L}\leq 1 for any L>0L>0 and if L=2πL=2\pi, γ2π=(arccos(1/3)2π)20.0925\gamma_{2\pi}=\left(\frac{\arccos(-1/3)}{2\pi}\right)^{2}\approx 0.0925.

Let us compute the constants needed to apply Proposition 2.1. As we do not expect an inequality of type (2.2) to hold in that case, we set KΣ,Ω:=+K_{\Sigma,\Omega}:=+\infty. Moreover, CΣC_{\Sigma} can be computed exactly as follows.

Lemma 3.12.

In this case, CΣ=1βγLC_{\Sigma}=\frac{1}{\beta\gamma_{L}}.

Proof.

The constant 1CΣ\frac{1}{C_{\Sigma}} is the smallest non-zero eigenvalue γ\gamma of the following problem:

{βΔτf=γfonΣ\{x0},fν=0at point xL,fe1+fe2+fe3=0at point x0,\displaystyle\begin{cases}\beta\Delta^{\tau}f=-\gamma f&\mbox{on}\ \ \Sigma\backslash\{x_{0}\},\\ \frac{\partial f}{\partial\nu}=0&\mbox{at point }x_{L},\\ \frac{\partial f}{\partial e_{1}}+\frac{\partial f}{\partial e_{2}}+\frac{\partial f}{\partial e_{3}}=0&\mbox{at point }x_{0},\end{cases}

where Δτ\Delta^{\tau} is the Laplace-Beltrami operator on Ω\partial\Omega and \mathcal{L}. A general solution to that boundary value problem is given by

f(x)={Acos(γβy)+Bsin(γβy)if x=(y,0),Ccos(γβθ)+Dsin(γβθ)if x=(cosθ,sinθ)Ω,\displaystyle f(x)=\begin{cases}A\cos(\sqrt{\frac{\gamma}{\beta}}y)+B\sin(\sqrt{\frac{\gamma}{\beta}}y)&\mbox{if }x=(y,0)\in\mathcal{L},\\ C\cos(\sqrt{\frac{\gamma}{\beta}}\theta)+D\sin(\sqrt{\frac{\gamma}{\beta}}\theta)&\mbox{if }x=(\cos\theta,\sin\theta)\in\partial\Omega,\end{cases}

where AA, BB, CC and DD have to satisfy the continuity assumption of ff at point x0x_{0} and both boundary conditions, that is:

{A=C=Ccos(γβ2π)+Dsin(γβ2π),0=Asin(γβL)+Bcos(γβL),0=B+D+Csin(γβ2π)Dcos(γβ2π).\displaystyle\left\{\begin{aligned} A&=C=C\cos(\sqrt{\frac{\gamma}{\beta}}2\pi)+D\sin(\sqrt{\frac{\gamma}{\beta}}2\pi),\\ 0&=-A\sin(\sqrt{\frac{\gamma}{\beta}}L)+B\cos(\sqrt{\frac{\gamma}{\beta}}L),\\ 0&=B+D+C\sin(\sqrt{\frac{\gamma}{\beta}}2\pi)-D\cos(\sqrt{\frac{\gamma}{\beta}}2\pi).\end{aligned}\right.

A short computation shows that this system has a non-trivial solution if and only if γβ\frac{\gamma}{\beta} solves (3.18). Therefore, 1CΣ=βγL\frac{1}{C_{\Sigma}}=\beta\gamma_{L}. Obviously, γ=1\gamma=1 is a solution to (3.18), thus γL1\gamma_{L}\leq 1. ∎

Next, we look for the constants K1K_{1} and K2K_{2}.

Lemma 3.13.

Inequality (2.3) holds with K1=38K_{1}=\frac{3}{8} and K2=L2(π+L)β(2π+L)K_{2}=\frac{L^{2}(\pi+L)}{\beta(2\pi+L)}.

Proof.

Recall that Σ=Ω\Sigma=\partial\Omega\cup\mathcal{L}. Let us insert the average of ff over Ω\partial\Omega as follows:

(Ωf𝑑λΩΣf𝑑λΣ)2\displaystyle\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2} 2(Ωf𝑑λΩΩf𝑑λ)2+2(Ωf𝑑λΣf𝑑λΣ)2\displaystyle\leq 2\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\partial\Omega}fd\lambda_{\partial}\right)^{2}+2\left(\int_{\partial\Omega}fd\lambda_{\partial}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2}
38Ωf2𝑑λΩ+2(Ωf𝑑λΣf𝑑λΣ)2,\displaystyle\leq\frac{3}{8}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}+2\left(\int_{\partial\Omega}fd\lambda_{\partial}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2},

where the second inequality follows directly from (3.3). Moreover, recalling that λΣ=2π2π+Lλ+L2π+Lλ\lambda_{\Sigma}=\frac{2\pi}{2\pi+L}\lambda_{\partial}+\frac{L}{2\pi+L}\lambda_{\mathcal{L}}

(Ωf𝑑λΣf𝑑λΣ)2\displaystyle\left(\int_{\partial\Omega}fd\lambda_{\partial}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2} =L2(2π+L)2(Ωf𝑑λf𝑑λ)2.\displaystyle=\frac{L^{2}}{(2\pi+L)^{2}}\left(\int_{\partial\Omega}fd\lambda_{\partial}-\int_{\mathcal{L}}fd\lambda_{\mathcal{L}}\right)^{2}.

For every x=(cosθ,sinθ)Ωx=(\cos\theta,\sin\theta)\in\partial\Omega, with θ(π,π]\theta\in(-\pi,\pi], we denote by pxp_{x} the point of \mathcal{L} with coordinates (1+L|θ|Lπ,0)(1+L-\frac{|\theta|L}{\pi},0). It follows that

(Ωf𝑑λf𝑑λ)2=(Ω(f(x)f(px))𝑑λ)2Ω(f(x)f(px))2𝑑λ.\displaystyle\left(\int_{\partial\Omega}fd\lambda_{\partial}-\int_{\mathcal{L}}fd\lambda_{\mathcal{L}}\right)^{2}=\left(\int_{\partial\Omega}(f(x)-f(p_{x}))d\lambda_{\partial}\right)^{2}\leq\int_{\partial\Omega}(f(x)-f(p_{x}))^{2}d\lambda_{\partial}.

Denoting by λ+\lambda_{\partial}^{+} and λ\lambda_{\partial}^{-} the normalized Hausdorff measures on Ω+:={(x,y)Ω:y>0}\partial\Omega^{+}:=\{(x,y)\in\partial\Omega:y>0\} and Ω:={(x,y)Ω:y<0}\partial\Omega^{-}:=\{(x,y)\in\partial\Omega:y<0\}, respectively,

Ω(f(x)f(px))2𝑑λ=12Ω+(f(x)f(px))2𝑑λ++12Ω(f(x)f(px))2𝑑λ.\displaystyle\int_{\partial\Omega}(f(x)-f(p_{x}))^{2}d\lambda_{\partial}=\frac{1}{2}\int_{\partial\Omega^{+}}(f(x)-f(p_{x}))^{2}d\lambda^{+}_{\partial}+\frac{1}{2}\int_{\partial\Omega^{-}}(f(x)-f(p_{x}))^{2}d\lambda^{-}_{\partial}.

Moreover, for any 𝒞1\mathcal{C}^{1}-function g:[π,L]g:[-\pi,L]\to\mathbb{R},

1π0π|g(θ)g(LθLπ)|2𝑑θπ+L2πL|g(t)|2𝑑t,\displaystyle\frac{1}{\pi}\int_{0}^{\pi}\left|g(-\theta)-g(L-\textstyle\frac{\theta L}{\pi})\right|^{2}d\theta\leq\frac{\pi+L}{2}\int_{-\pi}^{L}|g^{\prime}(t)|^{2}dt,

so we deduce, identifying Ω+\partial\Omega^{+} with [π,0][-\pi,0] and \mathcal{L} with [0,L][0,L], that

Ω+(f(x)f(px))2𝑑λ+π+L2(πΩ+τf2𝑑λ++Lτf2𝑑λ)\displaystyle\int_{\partial\Omega^{+}}(f(x)-f(p_{x}))^{2}d\lambda^{+}_{\partial}\leq\frac{\pi+L}{2}\left(\pi\int_{\partial\Omega^{+}}\|\nabla^{\tau}f\|^{2}d\lambda^{+}_{\partial}+L\int_{\mathcal{L}}\|\nabla^{\tau}f\|^{2}d\lambda_{\mathcal{L}}\right)

and using symmetry to deal with Ω\partial\Omega^{-}, we obtain

Ω(f(x)f(px))2𝑑λ\displaystyle\int_{\partial\Omega}(f(x)-f(p_{x}))^{2}d\lambda_{\partial} π+L4(πΩ+τf2𝑑λ++πΩτf2𝑑λ+2Lτf2𝑑λ)\displaystyle\leq\frac{\pi+L}{4}\left(\pi\int_{\partial\Omega^{+}}\|\nabla^{\tau}f\|^{2}d\lambda^{+}_{\partial}+\pi\int_{\partial\Omega^{-}}\|\nabla^{\tau}f\|^{2}d\lambda^{-}_{\partial}+2L\int_{\mathcal{L}}\|\nabla^{\tau}f\|^{2}d\lambda_{\mathcal{L}}\right)
(π+L)(2π+L)2Στf2𝑑λΣ.\displaystyle\leq\frac{(\pi+L)(2\pi+L)}{2}\int_{\Sigma}\|\nabla^{\tau}f\|^{2}d\lambda_{\Sigma}.

Putting together the above inequalities, we get

(Ωf𝑑λΩΣf𝑑λΣ)2\displaystyle\left(\int_{\Omega}fd\lambda_{\Omega}-\int_{\Sigma}fd\lambda_{\Sigma}\right)^{2} 38Ωf2𝑑λΩ+2L2(2π+L)2(π+L)(2π+L)2βΣβτf2𝑑λΣ\displaystyle\leq\frac{3}{8}\int_{\Omega}\|\nabla f\|^{2}d\lambda_{\Omega}+2\frac{L^{2}}{(2\pi+L)^{2}}\frac{(\pi+L)(2\pi+L)}{2\beta}\int_{\Sigma}\beta\|\nabla^{\tau}f\|^{2}d\lambda_{\Sigma}

which leads to inequality (2.3) with K1=38K_{1}=\frac{3}{8} and K2=L2(π+L)β(2π+L)K_{2}=\frac{L^{2}(\pi+L)}{\beta(2\pi+L)}. ∎

Proof of Proposition 3.11.

Since KΣ,Ω=K_{\Sigma,\Omega}=\infty, we immediately get from Proposition 2.1 that

Cαmax(CΩ+(1α)K1,αK2,CΣ+αK2)=max(CΩ+(1α)K1,CΣ+αK2).\displaystyle C_{\alpha}\leq\max\left(C_{\Omega}+(1-\alpha)K_{1},\alpha K_{2},C_{\Sigma}+\alpha K_{2}\right)=\max\left(C_{\Omega}+(1-\alpha)K_{1},C_{\Sigma}+\alpha K_{2}\right).

Therefore,

Cαmax(1σΩ+38(1α),1βγL+αL2(π+L)β(2π+L)),\displaystyle C_{\alpha}\leq\max\left(\frac{1}{\sigma_{\Omega}}+\frac{3}{8}(1-\alpha),\frac{1}{\beta\gamma_{L}}+\alpha\frac{L^{2}(\pi+L)}{\beta(2\pi+L)}\right), (3.19)

where σΩ3.39\sigma_{\Omega}\approx 3.39. ∎

Remark 3.14.

If β\beta is large enough, that is if the diffusion velocity is larger on Σ\Sigma than on Ω\Omega, then the first term in (3.19) dominates. Precisely, if βσΩ(1γL+L2(π+L)2π+L)\beta\geq\sigma_{\Omega}\left(\frac{1}{\gamma_{L}}+\frac{L^{2}(\pi+L)}{2\pi+L}\right), then (3.19) rewrites for any α\alpha

Cα1σΩ+38(1α).C_{\alpha}\leq\frac{1}{\sigma_{\Omega}}+\frac{3}{8}(1-\alpha).

Conversely, if β1γL(1σΩ+38)1\beta\leq\frac{1}{\gamma_{L}}\left(\frac{1}{\sigma_{\Omega}}+\frac{3}{8}\right)^{-1}, then (3.19) rewrites for any α\alpha

Cα1βγL+αL2(π+L)β(2π+L).C_{\alpha}\leq\frac{1}{\beta\gamma_{L}}+\alpha\frac{L^{2}(\pi+L)}{\beta(2\pi+L)}.

References

  • [1] Alexander Aurell and Boualem Djehiche, Behavior near walls in the mean-field approach to crowd dynamics, SIAM J. Appl. Math. 80 (2020), no. 3, 1153–1174. MR 4096131
  • [2] Dominique Bakry, Ivan Gentil, and Michel Ledoux, Analysis and geometry of Markov diffusion operators, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 348, Springer, Cham, 2014. MR 3155209
  • [3] William Beckner, Sharp Sobolev inequalities on the sphere and the Moser-Trudinger inequality, Ann. of Math. (2) 138 (1993), no. 1, 213–242. MR 1230930
  • [4] Rodney Josué Biezuner, Best constants in Sobolev trace inequalities, Nonlinear Anal. 54 (2003), no. 3, 575–589. MR 1978428
  • [5] Julián Fernández Bonder, Julio D. Rossi, and Raúl Ferreira, Uniform bounds for the best Sobolev trace constant, Adv. Nonlinear Stud. 3 (2003), no. 2, 181–192. MR 1971310
  • [6] Jean-Michel Bony, Philippe Courrège, and Pierre Priouret, Semi-groupes de Feller sur une variété à bord compacte et problèmes aux limites intégro-différentiels du second ordre donnant lieu au principe du maximum, Ann. Inst. Fourier (Grenoble) 18 (1968), no. fasc. 2, 369–521 (1969). MR 245085
  • [7] Jean-Dominique Deuschel, Giambattista Giacomin, and Lorenzo Zambotti, Scaling limits of equilibrium wetting models in (1+1)(1+1)-dimension, Probab. Theory Related Fields 132 (2005), no. 4, 471–500. MR 2198199
  • [8] Klaus-Jochen Engel, The Laplacian on C(Ω¯)C(\overline{\Omega}) with generalized Wentzell boundary conditions, Arch. Math. (Basel) 81 (2003), no. 5, 548–558. MR 2029716
  • [9] José F. Escobar, The geometry of the first non-zero Stekloff eigenvalue, J. Funct. Anal. 150 (1997), no. 2, 544–556. MR 1479552
  • [10] Torben Fattler, Martin Grothaus, and Robert Voßhall, Construction and analysis of a sticky reflected distorted Brownian motion, Ann. Inst. Henri Poincaré Probab. Stat. 52 (2016), no. 2, 735–762. MR 3498008
  • [11] Vincenzo Ferone, Carlo Nitsch, and Cristina Trombetti, On a conjectured reverse Faber-Krahn inequality for a Steklov-type Laplacian eigenvalue, Commun. Pure Appl. Anal. 14 (2015), no. 1, 63–82. MR 3299025
  • [12] Alexandre Girouard and Iosif Polterovich, Spectral geometry of the Steklov problem (survey article), J. Spectr. Theory 7 (2017), no. 2, 321–359. MR 3662010
  • [13] Gisèle Ruiz Goldstein, Derivation and physical interpretation of general boundary conditions, Adv. Differential Equations 11 (2006), no. 4, 457–480. MR 2215623
  • [14] Gisèle Ruiz Goldstein, Jerome A. Goldstein, Davide Guidetti, and Silvia Romanelli, Maximal regularity, analytic semigroups, and dynamic and general Wentzell boundary conditions with a diffusion term on the boundary, Ann. Mat. Pura Appl. (4) 199 (2020), no. 1, 127–146. MR 4065110
  • [15] Martin Grothaus and Robert Voßhall, Stochastic differential equations with sticky reflection and boundary diffusion, Electron. J. Probab. 22 (2017), Paper No. 7, 37. MR 3613700
  • [16] Nobuyuki Ikeda, On the construction of two-dimensional diffusion processes satisfying Wentzell’s boundary conditions and its application to boundary value problems, Mem. Coll. Sci. Univ. Kyoto Ser. A. Math. 33 (1960/61), 367–427. MR 126883
  • [17] James Kennedy, An isoperimetric inequality for the second eigenvalue of the Laplacian with Robin boundary conditions, Proc. Amer. Math. Soc. 137 (2009), no. 2, 627–633. MR 2448584
  • [18] Alexander V. Kolesnikov and Emanuel Milman, Brascamp-Lieb-type inequalities on weighted Riemannian manifolds with boundary, J. Geom. Anal. 27 (2017), no. 2, 1680–1702. MR 3625169
  • [19] Vitalii Konarovskyi and Max von Renesse, Reversible coalescing-fragmentating Wasserstein dynamics on the real line, 2017.
  • [20] Nikolay Kuznetsov and Alexander Nazarov, Sharp constants in the Poincaré, Steklov and related inequalities (a survey), Mathematika 61 (2015), no. 2, 328–344. MR 3343056
  • [21] Yanyan Li and Meijun Zhu, Sharp Sobolev trace inequalities on Riemannian manifolds with boundaries, Comm. Pure Appl. Math. 50 (1997), no. 5, 449–487. MR 1443055
  • [22] Svetlana Matculevich and Sergey Repin, Explicit constants in Poincaré-type inequalities for simplicial domains and application to a posteriori estimates, Comput. Methods Appl. Math. 16 (2016), no. 2, 277–298. MR 3483617
  • [23] Umberto Mosco, Composite media and asymptotic Dirichlet forms, J. Funct. Anal. 123 (1994), no. 2, 368–421. MR 1283033
  • [24] Delio Mugnolo, Robin Nittka, and Olaf Post, Norm convergence of sectorial operators on varying Hilbert spaces, Oper. Matrices 7 (2013), no. 4, 955–995. MR 3154581
  • [25] Delio Mugnolo and Silvia Romanelli, Dirichlet forms for general Wentzell boundary conditions, analytic semigroups, and cosine operator functions, Electron. J. Differential Equations (2006), No. 118, 20. MR 2255233
  • [26] A. I. Nazarov and S. I. Repin, Exact constants in Poincaré type inequalities for functions with zero mean boundary traces, Math. Methods Appl. Sci. 38 (2015), no. 15, 3195–3207. MR 3400329
  • [27] Andreas Nonnenmacher and Martin Grothaus, Overdamped limit of generalized stochastic hamiltonian systems for singular interaction potentials, 2018.
  • [28] Robert C. Reilly, Applications of the Hessian operator in a Riemannian manifold, Indiana Univ. Math. J. 26 (1977), no. 3, 459–472. MR 474149
  • [29] Julio D. Rossi, First variations of the best Sobolev trace constant with respect to the domain, Canad. Math. Bull. 51 (2008), no. 1, 140–145. MR 2384747
  • [30] Abdolhakim Shouman, Generalization of Philippin’s results for the first Robin eigenvalue and estimates for eigenvalues of the bi-drifting Laplacian, Ann. Global Anal. Geom. 55 (2019), no. 4, 805–817. MR 3951758
  • [31] M. A. Shubin, Pseudodifferential operators and spectral theory, second ed., Springer-Verlag, Berlin, 2001, Translated from the 1978 Russian original by Stig I. Andersson. MR 1852334
  • [32] Kazuaki Taira, Boundary value problems and Markov processes, third ed., Lecture Notes in Mathematics, vol. 1499, Springer, Cham, [2020] ©2020, Functional analysis methods for Markov processes. MR 4176673
  • [33] Satoshi Takanobu and Shinzo Watanabe, On the existence and uniqueness of diffusion processes with Wentzell’s boundary conditions, J. Math. Kyoto Univ. 28 (1988), no. 1, 71–80. MR 929208
  • [34] A. D. Ventcel, On boundary conditions for multi-dimensional diffusion processes, Theor. Probability Appl. 4 (1959), 164–177. MR 121855
  • [35] Hendrik Vogt and Jürgen Voigt, Wentzell boundary conditions in the context of Dirichlet forms, Adv. Differential Equations 8 (2003), no. 7, 821–842. MR 1988680
  • [36] Mahamadi Warma, The Robin and Wentzell-Robin Laplacians on Lipschitz domains, Semigroup Forum 73 (2006), no. 1, 10–30. MR 2277314
  • [37] Shinzo Watanabe, On stochastic differential equations for multi-dimensional diffusion processes with boundary conditions. II, J. Math. Kyoto Univ. 11 (1971), 545–551. MR 287612