Spectral gap estimates for Brownian motion on domains with sticky-reflecting boundary diffusion
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.
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 of a smooth domain with Feller generator
(1.1) | |||
where is the outer normal derivative, is the Laplace-Beltrami operator on the boundary and . The case of pure sticky reflection but no diffusion along the boundary corresponds to the regime ; models with 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 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 by Kennedy [17] and Shouman [30]. However, for the properties of the process change significantly, which is indicated by the fact that the energy form of 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 which corresponds to an inward sticky reflection at . Our ansatz to estimate the spectral gap is based on a simple interpolation idea. To this aim assume that and have finite (Hausdorff) measure so that we may choose for which
Introducing and as normalized volume and Hausdorff measures on and and setting
we find that is -symmetric with first nonzero eigenvalue/spectral gap characterized by the Rayleigh quotient
where
and
and denotes the tangential derivative operator on .
This representation of formally interpolates between the two extremal cases of the spectral gap for reflecting Brownian motion on when and for Brownian motion on the surface when . As our main result, in Proposition 2.1 we propose a simple method to estimate from below using only and and estimates for certain bulk-boundary interaction terms which are independent of . The method can lead to quite good results which is illustrated by the example when is a -dimensional unit ball. When and , for instance, it yields the estimate
where is the spectral gap for the Neumann Laplacian on the 2-dimensional unit ball, c.f. Section 3.1. – In case when is a -dimensional manifold with Ricci curvature bounded from below by and with boundary whose second fundamental form is bounded from below by we obtain (again with , for simplicity) that
where and are the usual (Neumann) Poincaré constants of and respectively. To derive this result we combine Escobar’s lower bound [9] on the first Steklov eigenvalue [12, 20] of with a novel estimate on the optimal zero mean trace Poincaré constant of [22, 26], for which we obtain that
for all with , 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 independent of the interpolation approach stating that
but which is generally weaker for small values of , c.f. Section 3.2.
The interpolation approach also yields a sufficient condition for the continuity of at , which in general may fail. In Section 2.2 we present sufficient conditions for continuity and discontinuity of at which hints towards a phase transition in the associated family of variational problems.
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 be an open domain in or a Riemannian manifold with a piecewise smooth boundary . Let be a smooth compact and connected subset of . We denote by the boundary of in the space , i.e. . We consider two probability measures and with support and , which are absolutely continuous with respect to the Lebesgue and the Hausdorff measure on and , respectively.
Let and denote given first order gradient operators mapping differentiable functions into (tangential) vector fields on and on , respectively, and for let
where is dense in . We assume that for the quadratic form is a pre-Dirichlet form on whose closure we shall denote by , c.f. [15] for details. We wish to estimate from above , where is the optimal Poincaré constant given by
(2.1) |
In the interpolation method presented below it is assumed that are known or can be estimated at the two extremals . For instance, when , are the standard gradient resp. tangential gradient operators and and are normalized Lebesgue resp. Hausdorff measures on and , is the optimal Poincaré constant associated to the Laplace operator on with Neumann boundary conditions, whereas is the optimal Poincaré constant associated to the Laplace-Beltrami operator on with Neumann boundary conditions on .
The following proposition establishes an estimate of in terms of and .
Proposition 2.1.
Assume there exists constants , such that for any
(2.2) |
and
(2.3) |
then it holds for any ,
(2.4) |
Proof.
By definition of and by (2.2), for any
for any . Let . For any and any
Therefore,
For any positive constants , we have
Therefore
The last term is equivalent to the announced result. ∎
2.2 Continuity of
In general, the function might have discontinuities at in which cases an upper bound for which interpolates continuously between and cannot exist. For example, when and , straightforward computations yield
where . Hence is discontinuous at if and only if . – To generalize this to the framework of Section 2.1 let and
(If , is the inverse of the spectral gap for Brownian motion on with killing on and normal reflection at . ) 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
In particular, if , then is discontinuous at . Conversely, if then is continuous at 0. If continuity at 1 holds.
Proof.
To prove the second statement, take a non constant function and estimate
Since on , we obtain
Taking the supremum over yields the first statement.
Remark 2.3.
For smooth enough boundary the constant can always be taken equal to zero, hence by proposition 2.2 continuity at holds. An example where a phase transition appears at is given in section 3.3. In section 3.4 we present an example where but continuity of at 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 be the unit ball in , and and with .
Proposition 3.1.
In the case when the optimal Poincaré constant of the generator (1.1) is bounded from above by
(3.1) |
where and is the optimal Poincaré constant for reflecting Brownian motion on .
Proof.
In order to apply Proposition 2.1, it is sufficient to compute the constants , , and . We claim that inequalities (2.2) and (2.3) holds with
First, according to [31, Theorem 22.1], the first eigenvalue of the Laplace-Beltrami operator on the unit sphere of dimension is equal to , thus .
Moreover, according to [3, Theorem 4], for every one has
for if and if , where is the harmonic extension of to the unit ball . It implies the logarithmic Sobolev inequality . Repeating the proof of Proposition 5.1.3 in [2], we get . Moreover, since the harmonic extension of is minimizing the energy functional under any function with boundary condition , the last inequality implies for any
(3.2) |
which implies .
Furthermore, note that , where , . Hence, using Jensen’s inequality and polar coordinates
We separately estimate
for any . Hence,
(3.3) |
which implies and . ∎
For illustration, in we compare the bound from Proposition 3.1 for to the optimal constant which will be computed numerically. To evaluate the bound (3.1), note that in this case
(3.4) |
where 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 , , where is the Bessel function of the first kind of parameter , defined by , . As a consequence, inequality (3.1) becomes
(3.5) |
For the numerical computation of one notes that the generator associated with is defined on as
where and denote the Laplace-Beltrami operator and the outer normal derivative on the circle . Hence, an eigenvector of for eigenvalue is a function such that
This equation is equivalent to the system of partial differential equations
which by the continuity of can be rewritten as
Passing to polar coordinates in and separating variables, we obtain the set of eigenfunctions ,
where , , are countable family of positive solutions to the equation
(3.6) |
for every . Since the family is dense in and the operator is symmetric, the standard argument implies
(3.7) |
where . The resulting curves are plotted in Figure 1.

3.2 Smooth manifold with boundary
Let be a smooth compact Riemannian manifold of dimension with piecewise smooth boundary . We denote by the Ricci curvature of and by the second fundamental form on the boundary . Assume in this section that:
As before we consider , and with .
Proposition 3.2.
Under assumption (M), it holds that
(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 and
Proof.
Our goal is to obtain an lower bound of
where we recall that . We note that
Let be a minimizer for . Then and
for each with . By integration by parts, the latter equality is equivalent to
for each satisfying . In particular, choosing (which obviously satisfies ), we get that should satisfy in . Hence for each with zero mean, so it follows that for every , which is equivalent to
for every . It follows that is constant on . Therefore, satisfies
(3.9) |
for some constant .
Moreover, recall Reilly’s formula (see [28])
(3.10) | ||||
where and denote the Riemannian volume resp. surface measure on and , is the Hessian of and is the mean curvature of (i.e. the trace of ). Since satisfies (3.9),
because . Furthermore, note that . Therefore, the l.h.s. of (3.10) is bounded by
On the other hand, by assumption (M), , and
Since
the r.h.s. of (3.10) is bounded from below by . It turns out that
which implies that . It follows that inequality (2.3) holds with . ∎
Remark 3.4.
Instead of using from Proposition 3.3 another admissible choice is
where is the optimal Sobolev trace constant of , i.e. the norm of the embedding . is the first nontrivial eigenvalue of a Steklov-type eigenvalue problem
for which however explicit lower bounds in terms of the geometry of seem yet unknown [4, 5, 11, 21, 29].
Proposition 3.5.
Under assumption (M), inequality (2.2) holds with .
Proof.
Alternatively, we obtain another upper bound for by a direct application of Reilly’s formula.
Proposition 3.6.
Under assumption (M) it holds that
(3.11) |
Proof.
We estimate equivalently from below the first nontrivial eigenvalue for the problem
where . As in the proof of Proposition 3.3 we apply Reilly’s formula (3.10) to the corresponding eigenfunction . In this case, for the l.h.s. we estimae
Since
the r.h.s. of (3.10) is bounded from below by
Combining the two bounds for (3.10) yields
which implies that either
or
Consequently,
∎
Corollary 3.7.
When goes to 0, tends to and tends to , so the estimation via the interpolation method is always stronger. When goes to , tends to and tends to , so the relative strength of each method depends on the values of , and .
3.3 Brownian motion on balls with partial sticky reflecting boundary diffusion
As in Section 3.1, let be the unit ball of . Now, define for a fixed
Proposition 3.8.
It holds that
(3.12) |
where and .
As previously, we will start by computing the needed constants , , , and . The first constant, , remains unchanged.
Lemma 3.9.
The following inequalities hold true
(3.13) | ||||
(3.14) |
where and .
Proof.
Lemma 3.10.
It holds that
with .
Proof.
For every with polar coordinates , , , denote by the point of coordinates on . Obviously, and by Jensen’s inequality
Define . Then
(3.15) |
where and . On the one hand
(3.16) |
On the other hand
For every , , thus
(3.17) |
The proof of the lemma is completed by putting together (3.15), (3.16) and (3.17). ∎


For sufficiently large, the map is continuous at . Indeed, by Proposition 2.2, a sufficient condition is , that is
which is satisfied for any .
3.4 Ball with a needle
Our final example is the unit ball of with a needle of length attached to one point of the boundary, i.e. , see Figure 3. The attachment point and the endpoint of the needle are denoted by and , respectively.

In that setting, we define , and
where is as previously the normalized Lebesgue measure on and , with and being the normalized Hausdorff measures on and , respectively. We choose
where , and are the three ”tangent” vectors to at point , and , , which is well defined in . With this choice, for is a pre-Dirichlet form on , whose closure generates Brownian motion on with sticky boundary diffusion on , i.e. whose generator is given by
with being the generator of the canonical diffusion on with reflecting boundary condition at . As before, the optimal Poincaré constant for is given by
and let and . In this case the following estimate is obtained.
Proposition 3.11.
where is the smallest positive solution to
(3.18) |
Note that for any and if , .
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 . Moreover, can be computed exactly as follows.
Lemma 3.12.
In this case, .
Proof.
The constant is the smallest non-zero eigenvalue of the following problem:
where is the Laplace-Beltrami operator on and . A general solution to that boundary value problem is given by
where , , and have to satisfy the continuity assumption of at point and both boundary conditions, that is:
A short computation shows that this system has a non-trivial solution if and only if solves (3.18). Therefore, . Obviously, is a solution to (3.18), thus . ∎
Next, we look for the constants and .
Lemma 3.13.
Inequality (2.3) holds with and .
Proof.
Recall that . Let us insert the average of over as follows:
where the second inequality follows directly from (3.3). Moreover, recalling that
For every , with , we denote by the point of with coordinates . It follows that
Denoting by and the normalized Hausdorff measures on and , respectively,
Moreover, for any -function ,
so we deduce, identifying with and with , that
and using symmetry to deal with , we obtain
Putting together the above inequalities, we get
which leads to inequality (2.3) with and . ∎
Proof of Proposition 3.11.
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 -dimension, Probab. Theory Related Fields 132 (2005), no. 4, 471–500. MR 2198199
- [8] Klaus-Jochen Engel, The Laplacian on 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