1. Introduction
Let be a complete Riemannian manifold of dimension . For each compact domain with smooth boundary, let denote the first eigenvalue of the Laplacian operator on with Dirichlet boundary condition. Given , a very interesting isoperimetric probem in spectral geometry and shape optimization is that of investigating the minimizers or, more generally, the critical points of the functional defined on the class of domains such that .
Regarding to this problem, Faber [10] and Krahn [18] proved that, in , geodesic balls are the only global minimizers of subject to a volume constraint. More precisely, they proved that if is a compact domain, then , where is a geodesic ball such that , and equality holds if and only if is a geodesic ball. We note that the analogous result is true in , the round sphere, and , the hyperbolic space (see, for example, Chapter IV, Section 2, of [6]).
In turn, it is known that a compact domain with smooth boundary is a critical point of the functional with respect to volume preserving deformations of in if and only if the first eigenfunctions of the Laplacian operator of with Dirichlet boundary condition solve the following overdetermined problem:
|
|
|
where is a constant and denotes the outward unit normal vector along . This follows from the first variation formula for the first Dirichlet eigenvalue of the Laplacian operator, known as Hadamard formula, which was originally proved by Garabedian and Schiffer [12] in Euclidean space and later by El Soulfi and Ilias [8] to general Riemannian manifolds. From now on, following [22], we say that a compact domain with smooth boundary is an extremal domain if it is a critical point of with respect to volume preserving deformations. We refer the reader to Section 2 for details and precise definitions.
In Euclidean space , in hyperbolic space , and in the round hemisphere , only geodesic balls are extremal domains. This result follows from the classical work of Serrin [27], and its extension to space forms by Kumaresan and Prajapat [19], on elliptic overdetermined problems, which was proved using the moving plane method introduced by Alexandrov in [2].
Extremal domains share many similarities with embedded hypersurfaces of constant mean curvature, and significant progress has been made in recent years regarding their existence, regularity, and classification. For example, in [22], Pacard and Sicbaldi demonstrated the existence of extremal domains with small volume in compact Riemannian manifolds by assuming that the scalar curvature of has a nondegenerate critical point. In [7], Delay and Sicbaldi improved this result by removing the condition on the scalar curvature of .
Recently, Lamboley and Sicbaldi [20] established that for any connected compact Riemannian manifold and given , there exists an open set that is smooth except for a singular set of codimension less than , satisfying
|
|
|
In particular, is a smooth extremal domain if or . On the other hand, Espinar and Mazet [9] proved that simply connected extremal domains (and, in fact, -extremal disks) in are geodesic disks.
It is worth mentioning that all the above results have parallels in the context of hypersurfaces of constant mean curvature (see, for example, [31], [21], [3], [15], [13], and [17]). For more results on the existence of extremal domains, we refer the reader to [28] and [29].
Furthermore, numerous examples of extremal domains can be constructed on Riemannian manifolds endowed with an isometric action by a compact Lie group (see Section 2.1). In particular, this construction provides many examples of extremal domains on the round sphere that are not geodesic balls. This shows that the moving plane method used by Serrin in [27] fails in this case.
In this paper, we initiate the investigation of the stability properties of extremal domains. We say that an extremal domain is stable if it minimizes the first Dirichlet eigenvalue functional up to the second order with respect to volume preserving deformations.
In Section 3, we derive the second variation formula and establish a criterion for stability, obtaining important insights into the geometry of extremal domains. Using this criterion, we give a complete characterization of stable extremal domains in the round sphere . We prove the following theorem:
Theorem 1.1.
Let be an extremal domain. If is stable, then it is necessarily a geodesic disk.
This theorem can be considered as the parallel of the Barbosa-do Carmo-Eschenburg theorem [4]. It is interesting to observe that Theorem 1.1, combined with Lamboley-Sicbaldi’s theorem proved in [20] and cited above, provides an alternative proof of the Faber-Krahn inequality in , which states that for any , if is a compact domain with area , then:
|
|
|
where is a geodesic disk with area and equality holds if and only if is a geodesic disk.
In the case that and , Theorems 1.1 and 1.2 in [20] imply that there exists a smooth domain such that . In particular, is a stable extremal domain and, in fact, a global minimizer of with an area constraint. By Theorem 1.1 above, we have that is a geodesic disk . Moreover, if , then is a stable extremal domain. Applying Theorem 1.1 again we have that is also a geodesic disk, thus proving the Faber-Krahn inequality in .
In general, for stable extremal domains in a given Riemannian surface , we prove the following restriction on the topology of .
Theorem 1.2.
Let be an orientable compact Riemannian surface. If is a stable extremal domain with nonnegative total Gauss curvature, that is, , then the only possible values for the genus of and the number of connected components of are:
-
(1)
and ;
-
(2)
and .
In particular, it follows that stable extremal domains in a sphere with nonnegative curvature have at most boundary components.
Since the Faber-Krahn inequality is true on for any , it is a very natural question to ask if the characterization proved in Theorem 1.1 can be extended to stable extremal domains in , for . By using a result due to Reilly (see [23, Theorem 4]), we are able to give the following description of stable extremal domains , , in the case is a minimal hypersurface.
Theorem 1.3.
Let be an extremal domain such that is a compact minimal hypersurface. If is stable, then is a hemisphere.
Finally, we would like to mention that our methods are inspired by those used by Ros and Vergasta [25] to study stable constant mean curvature hypersurfaces with free boundary in convex domains of .
2. Basic definitions and first variation formula for
Let be a smooth Riemannian manifold. For each smooth compact domain , let denote its first eigenvalue of the Laplacian operator with Dirichlet boundary condition. We will start this section by defining what we mean by a local deformation of in .
A smooth local deformation of in is a smooth one-parameter family of domains , , given by , where is the smooth flow of some smooth vector field . If, in addition, for all , then we say that is a volume preserving local deformation of .
Let , , be a smooth local deformation of a smooth compact domain . As the first Dirichlet eigenvalue of the Laplacian operator is simple, it follows from the implicit function theorem that the function is smooth.
Definition 2.1 (Extremal domains).
We say that a smooth compact domain
is extremal if
|
|
|
for any volume preserving local deformation , , of in .
Next, we will state the Hadarmad formula proved by El Soufi and Ilias [8] which gives a first variation formula for the functional .
Proposition 2.2 (El Soufi and Ilias, [8]).
Let be a smooth Riemannian manifold and let be a smooth compact domain. Let , , be a smooth local deformation of . We have
(2.1) |
|
|
|
where is the outward unit normal vector along ,
is the normal displacement of induced by the deformation,and
is the first Dirichlet positive eigenfunction of the Laplacian operator on such that
We notice that if , , is volume preserving, that is, for all , then . Conversely, if is such that , then there exists a smooth local deformation , , such that on .
Therefore, it follows as a consequence of (2.1), that is an extremal domain if and only if its first eigenfunctions of Laplacian operator with Dirichlet boundary condition solve the following overdetermined elliptic problem:
|
|
|
where .
2.1. Examples
Assume that there exists an isometric action of a compact Lie group on a Riemannian manifold , and let be a domain invariant under the action of with a single connected boundary component. If is a first eigenfunction, then also solves the Dirichlet eigenvalue problem in for all . Since is simple, it follows that , defining a group homomorphism . Due to the compactness of , we have , proving that is -invariant. This invariance implies that is constant along .
A basic non-trivial example using this construction is given by the action on the unit round sphere when we consider . In this case, the invariant domains are the solid tori , , whose boundaries are given by the principal orbits of , . In particular, is a minimal hypersurface.
Note that, by choosing suitable values for , we can also construct extremal -domains from this family with two boundary components.
Another family of examples of extremal domains we can construct using this method are the solid tori , , where stands for the flat -torus. From this particular family, Sicbaldi [28] and Schlenk-Sicbaldi [26] constructed new extremal domains using bifurcation techniques.
3. Second variation formula for and stability of extremal domains
Let and be a smooth Riemannian manifold and a smooth compact domain, respectively. Suppose that is extremal. Since is a critical point of the functional , it is natural to investigate the second variation of this functional at . In this section, we will present a second variation formula for this functional, in a very general setting. As a consequence, we will give a criterion for a given extremal domain to be stable.
Definition 3.1 (Stable extremal domains).
We say that an extremal domain is stable if
|
|
|
for any volume preserving local deformation , , of in .
In the following proposition, we present a formula for the second variation for the functional at an extremal domain , with respect to volume preserving local deformations.
Proposition 3.2 (Second variation formula for ).
Let be an extremal domain and let be the first Dirichlet positive eigenfunction of the Laplacian operator on with . For any volume preserving local deformation , , of in , we have
(3.1) |
|
|
|
where is the mean curvature of with respect to the outward unit normal vector , is the normal displacement of induced by the deformation, and is a -extension of , that is, is a solution to the following problem:
|
|
|
We recall that volume preserving local deformations of are in bijection with functions in the space of smooth functions on the boundary of with zero average. Motivated by Proposition 3.1, we consider the quadratic form
|
|
|
given by
|
|
|
where is a -extension of to .
This setup allows us to define the index of an extremal domain.
Definition 3.5 (Index of Extremal Domains).
If is an extremal domain, then the Morse index of is given by
|
|
|
By this definition, the index represents the number of linearly independent, volume-preserving variations that decrease the first Dirichlet eigenvalue. In particular, is stable if and only if .
Now, we have the following useful criterion for the stability of extremal domains.
Proposition 3.6.
An extremal domain is stable if and only if
|
|
|
for all such that .
Proof.
Given , define
|
|
|
where .
Integrating by parts, we obtain that
(3.3) |
|
|
|
where is a -extension of .
Now, since for all such that on , we have that
(3.4) |
|
|
|
for all . Therefore, from (3.3) and (3.4) we conclude the proof.
∎
We conclude this section with some properties of Jacobi functions associated to the first Dirichlet eigenvalue.
We say that a function is a Jacobi function of if
|
|
|
for all .
The next lemma presents basic facts about Jacobi functions of that will be used in Section 5.
Lemma 3.7.
Let be an extremal domain.
-
(i)
is a Jacobi function of if and only if there exist and a -extension of such that
|
|
|
-
(ii)
If is stable and satisfies
|
|
|
then is a Jacobi function.
5. Proof of Theorem 1.1
Let be a stable extremal domain and let , , be the connected components of .
For each unit vector , define the following vector field on : , where denotes the vector product in .
Now, define by , where denotes the outward unit normal vector along . Since , we have
(5.1) |
|
|
|
for all .
The vector field generates a one-parameter family of rotations around the vector . Consider the local deformation of in given by . As is an isometry of for all , we have that for all . In particular,
|
|
|
and, by Proposition 3.2, this implies that
|
|
|
Therefore, since is stable, we have by Lemma 3.7 that is a Jacobi function of , that is, there exists a function such that
(5.2) |
|
|
|
After adding to a constant multiple of , if necessary, where is the first eigenfuncion of the Laplacian operator on with Dirichlet boundary condition such that and , we may assume without loss of generality that in the second equation above.
Next, fix . We know that consists of two connected components, both homeomorphic to a disk. Let be one of these connected components. Note that . Now, let be the closed geodesic disk of largest radius such that . Because of the way was chosen, we have that touches tangentially at least at two points. In particular, if is the center of , then vanishes at least at two points on . We claim that vanishes at least at one more point of . In fact, if , we are done. If not, we let be the connected component of which contains and consider , the geodesic closed disk centered at contained in with the largest possible radius. In this case, we have that touches tangentially at least at one point. Thus, this proves our claim.
Following ideas from Ros and Vergasta [25], we will prove in the following that has at least three connected components. Let be the connected components of . By applying the Gauss-Bonnet Theorem to each we obtain
|
|
|
where are the external angles of .
By summing over , we get that
|
|
|
where stands for the sum of the external angles of all nodal domains .
By the Gauss-Bonnet Theorem applied to , we have
(5.3) |
|
|
|
Now, it follows from (5.1) that vanishes in at least two points of , for all . Moreover, we know that for vanishes in at least three points of . This implies that
(5.4) |
|
|
|
Thus, by (5.3) and (5.4) we can conclude that
|
|
|
This implies that has at least three connected components.
But since is a stable extremal domain and solves (5.2) we have that has at most two connected components unless . Thus, it follows from the above argument that , which implies that is rotationally symmetric.
Since the rotationally symmetric extremal annuli are unstable we have that is a geodesic ball.
6. Proof of Theorem 1.2
Let be an orientable Riemannian surface and let be a stable extremal domain. Suppose that has nonnegative total Gauss curvature, that is, .
By a result of Gabard [11, Théorème 7.2.], which improved a previous result due to Alfhors [1], there exists a proper conformal branched cover , where is the closed unit disk, of degree at most . Moreover, after composing with a conformal diffeomorphism of if necessary, we may suppose that
|
|
|
for .
Thus, by the stability of , we have that
|
|
|
Since is conformal and has degree at most ,
|
|
|
Thus, it follows from this and from the Gauss-Bonnet Theorem that
|
|
|
Thus, , that is, .
Now, in order to estimate the number of connected components of , we use a similar, but different, balancing argument.
Let denote a compact Riemannian surface obtained from by attaching a conformal disk at any connected component of , such that
and have the same genus.
There exists a nonconstant holomorphic map of degree at most (see, for example, [14], p.261), where is the unit sphere centered at the origin.
Again, after composing with a conformal diffeomorphism of if necessary, we may assume that
|
|
|
for .
For each , we define
. Since in , it follows from the stability of that
|
|
|
|
|
|
|
|
|
|
|
|
By the Gauss-Bonnet Theorem, we have that
|
|
|
Therefore, if , then which implies that . In the case , we have and this implies .
7. Proof of Theorem 1.3
Suppose that is a stable extremal domain such that is a compact minimal hypersurface, that is, . It follows from a result due to Reilly [24, Theorem 4] that with equality if and only if is a hemisphere.
Since is minimal, it is well known that
|
|
|
for all .
Thus, using the coordinate functions as test functions in Proposition 3.6, we get
|
|
|
for all .
By summing over and using that , we have
|
|
|
This implies . Thus, and is a hemisphere.
Appendix A Proof of Proposition 3.2
Let be a Riemannian manifold and let be an extremal domain. In this section, we will deduce the second variation formula (3.1) for the first eigenvalue functional at .
Let be the first eigenfunction of the Laplacian operator on with Dirichlet boundary condition such that on and . Since is extremal, there is a constant such that on .
For each smooth domain , define
|
|
|
Let , , be a local deformation of in given by a smooth vector field . It is well known that
|
|
|
where , denotes the outward unit normal vector along , and is the volume element of induced by .
Therefore, by Proposition 2.2, we have
(A.1) |
|
|
|
where denotes the positive first eigenfunction of the Laplacian operator on with Dirichlet boundary condition and such that .
Since , it follows from (A.1) that is a critical point of with respect to all local deformations of in , not only the volume-preserving ones. Moreover, we have
|
|
|
|
|
|
|
|
From now on, we will denote derivatives with respect to at by using dots.
Note that
|
|
|
|
|
|
|
|
|
|
|
|
Let be the pullback metric of on . Let be the derivative tensor at . Note that , where denotes the Lie derivative.
Since on , we get . Moreover, since for all , it follows directly that
|
|
|
Therefore,
|
|
|
|
|
|
|
|
Now, suppose that , , is a volume-preserving local deformation. Note that, in this case, we have
|
|
|
Moreover, we also have that .
Let be a -extension of , that is, solves the following problem (see Remark 3.3):
|
|
|
Using that along , the Green formula yields
|
|
|
|
|
|
|
|
where is the volume element of .
Now, differentiating the equation we get
|
|
|
And so, since we have
|
|
|
Combining these identities, we arrive at
(A.2) |
|
|
|
where (see [5, 8]).
Integrating by parts, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Plugging this identity into (A.2) and noting that
|
|
|
we get
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Using , we have
|
|
|
|
|
|
|
|
On the other hand,
|
|
|
|
|
|
|
|
|
|
|
|
and similarly,
|
|
|
|
Thus, we have that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Now, on , we have and , where is the component of tangent to . Moreover, note that
|
|
|
Thus
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Next, let us deal with the last integral in the formula of by applying integration by parts again and using that .
We have:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where in the last equality we have used that . Now, since , we have that
|
|
|
|
|
|
|
|
By using that , we obtain that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
As on , we obtain that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Thus, substituting into the formula above for and canceling some terms, we have that
|
|
|
|
|
|
|
|
|
|
|
|
and this finishes the proof.
Acknowledgments
M.C. would like to express his gratitude to Princeton University for its hospitality, especially to Fernando Codá Marques, whose insightful questions inspired this work. This work was carried out while the authors were visiting ICTP - International Centre for Theoretical Physics, as associates. M.C. and I.N. acknowledge support from the ICTP through the Associates Programme (2022-2027 and 2019-2024, respectively). The authors are very grateful to the Institute and to Claudio Arezzo for their hospitality.
The authors were partially supported by the Brazilian National Council for Scientific and Technological Development (CNPq) under Grant 405468/ 2021-0. M.C. was also supported by CNPq under Grant 11136/ 2023-0 and by the Foundation for Research Support of the State of Alagoas (FAPEAL) under Grant 60030.0000000323/2023.