Low index capillary minimal surfaces
in Riemannian -manifolds
Eduardo Longa
Departamento de Matemática, Instituto de Matemática e Estatística, Universidade de São Paulo, R. do Matão 1010, São Paulo, SP 05508-900, Brazil
[email protected]
Abstract.
We prove a local rigidity result for infinitesimally rigid capillary surfaces in some Riemannian -manifolds with mean convex boundary. We also derive bounds on the genus, number of boundary components and area of any compact two-sided capillary minimal surface with low index under certain assumptions on the curvature of the ambient manifold and of its boundary.
††The author was partially supported by grant 2017/22704-0, São Paulo Research Foundation (FAPESP).
1. Introduction
Minimal surfaces are critical points for the area functional under suitable constraints. If we only consider closed surfaces, no constraints are necessary. If the surfaces have a fixed boundary, this leads to the so called Plateau problem, first studied by Lagrange [11] and Meusnier [14]. However, there is a third situation that can be investigated: when we consider compact surfaces whose boundaries are allowed to move freely in the boundary of the ambient manifold. Not surprisingly, critical points of the area functional under this constraint are called free boundary minimal surfaces. The boundary of such surfaces meet the boundary of the ambient manifold at a angle.
Although the first works dealing with this type of surfaces date back to 1938 with R. Courant (see [8] and [9]), in the last decade there have been incredible developments in this field, with the employment of new techniques and the emergence of interesting conceptual links. Among the main contributors to this topic of research we could cite Fraser, Chen and Pang, with their work on free boundary surfaces on positively curved ambients [6], Ambrozio with his work on rigidity of mean-convex manifolds [2], and additionally Carlotto and Sharp (jointly with Ambrozio), with their works on compactness analysis and index estimates for free boundary minimal hypersurfaces ([3] and [4]).
In this paper we are interested in a natural generalisation of free boundary minimal surfaces, namely, capillary minimal surfaces. These are critical points of a certain energy functional, which will be presented in Section 2. As will be deduced later, they can be characterised as minimal surfaces whose boundary meet the ambient boundary at a constant angle — the capillary angle.
Capillary surfaces in model the configuration of liquids in containers in the absence of gravity. In fact, the interface between the fluid and the air is a surface with boundary that (locally) minimises the energy functional. This energy depends on the area of the interface, the area wetted by the fluid in the container and the angle of contact between the surface and the boundary of the container. More general situations have also been considered in the literature, like the influence of gravity and the density of the fluid in the equilibrium shape. We refer the interested reader to the book of Finn [10] for an extensive survey on this subject and the derivation of the equations that describe such surfaces.
Like in the free boundary case, questions relating the topology and the geometry of surfaces raise a lot of attention from geometers. For instance, given an ambient manifold of a particular shape, what are the possible topological types of surfaces that admit a capillary CMC or minimal embedding into ? Is it possible to characterise the geometry of the allowed types? The first result in this direction was obtained by Nitsche [15], who proved that any immersed capillary disc in the unit ball of must be either a spherical cap or a flat disc. Later, Ros and Souam [18] extended this result to capillary discs in balls of -dimensional space forms. Recently, Wang and Xia [21] analysed the problem in an arbitrary dimension and proved that any stable immersed capillary hypersurface in a ball in space forms is totally umbilical. There are many interesting uniqueness results in other types of domains, like slabs [1], wedges [16, 7], cylinders [13] and cones [17].
We are interested in the same questions raised above, but in a more general setting. Namely, we only impose curvature assumptions on the ambient -manifold and look for restrictions in the topology of the possible immersed (or embedded) capillary minimal surfaces.
When dealing with the free boundary case, Ambrozio introduced the following functional in the space of
compact and properly immersed surfaces in a Riemannian -manifold :
where is the scalar curvature of , is the mean curvature of , denotes the area of and denotes the length of .
As a first result, we modify Ambrozio’s functional to take care of the capillary case and we show the following theorem, generalising Proposition 6 in [2]:
Theorem A.
Let be a Riemannian -manifold with nonempty boundary, and assume that and are bounded from below. If is a compact two sided capillary stable minimal surface, immersed in with contact angle , then
where denotes the Euler characteristic of . Moreover, equality occurs if and only if satisfies the following properties:
(i)
is totally geodesic in and the geodesic curvature of in is equal to ;
(ii)
the scalar curvature is constant along and equal to , and the mean curvature is constant along and equal to ;
(iii)
and , where is a unit normal for and denotes the second fundamental form of in the direction of a unit conormal for in .
In particular, (i), (ii) and (iii) imply that has constant Gaussian curvature equal to and has constant geodesic curvature equal to .
A compact two-sided capillary minimal surface, properly embedded in with contact angle , that satisfies conditions (i), (ii) and (iii) of Theorem A will be called infinitesimally rigid. Given one such surface , there is a way to obtain a vector field in such that and is tangent to along . Let us also denote by an extension to which is tangent to along the entire boundary of . Let the local flow of and fix a number between and . We show the existence of a local foliation around , employing the same techniques as Ambrozio in Proposition 10 in [2]:
Theorem B.
Let be a Riemannian -manifold with nonempty boundary, and assume that and are bounded from below. Let be a compact two-sided capillary minimal surface, properly embedded in with contact angle . If is infinitesimally rigid, then there exists and a map such that for every , the set
is a capillary CMC surface with contact angle and mean curvature . Moreover, for each and ,
In particular, taking a smaller if necessary, is a capillary CMC foliation of a neighbourhood of in .
It is also possible to show the existence of a capillary minimal foliation, where the contact angles now vary from leaf to leaf. Although it will not be used subsequently, we believe it may be of independent interest (see Remark 3.4).
Theorem C.
Let be a Riemannian -manifold with nonempty boundary, and assume that and are bounded from below. Let be a compact two-sided capillary minimal surface, properly embedded in with contact angle . If is infinitesimally rigid, then there exists and a map such that for every , the set
is a capillary minimal surface with contact angle . Moreover, for each and ,
In particular, taking a smaller if necessary, is a capillary minimal foliation of a neighbourhood of in .
Then, we use Theorem B to show that, under some hypotheses, a dichotomy occurs: either the contact angle is equal to or a very special situation takes place. More precisely, we have:
Theorem D.
Let be a Riemannian -manifold with nonempty and weakly mean-convex boundary, and assume that is bounded from below. Let be an energy-minimising and infinitesimally rigid surface, properly embedded in with contact angle . Assume that one of the following hypothesis holds:
(a)
each component of is locally length-minimising in ; or
(b)
.
Then either or is a flat and totally geodesic cylinder, is flat and is totally geodesic around . In the first case, there is a neighbourhood of in that is isometric to , where has constant Gaussian curvature and has constant geodesic curvature in .
A situation when may happen in Theorem D is the following. Let and be two non parallel planes in , intersecting along a line , and let be a plane in which is parallel to and intersects both and at the same angle. Fix to be the (closed) wedge determined by and which contains both and , and let (see Figure 1). Now fix a translation of by a vector parallel to the line and let be the quotient of by the group generated by . If we define to be the quotient of by , then is an infinitesimally rigid cylinder intersecting at a constant angle, (a) and (b) hold, is flat and is totally geodesic, as we wanted. One question remains: is energy-minimising? We believe so, but we did not find a proof.
Figure 1. An arrangement of planes in seen from a plane orthogonal to the line .
Next, we prove two other inequalities relating the geometry and the topology of capillary minimal surfaces of low index. This result generalises Theorem 1.2 in [6] to the capillary case. We note that item (i) below also generalises the free boundary case.
Theorem E.
Let be a Riemannian -manifold with nonempty boundary. Suppose that is a compact orientable two-sided capillary minimal surface of genus and with boundary components, immersed in with contact angle .
(i)
Suppose that has nonnegative Ricci curvature and weakly mean-convex boundary. If has index then
In particular, if the total geodesic curvature of (in ) is nonnegative (which happens if is free boundary and is weakly convex, for instance), then
(a)
if is even;
(b)
if is odd.
(ii)
Suppose that the scalar curvature of and the mean curvature of are bounded from below. If has index , then
(iii)
Suppose that has scalar curvature and weakly mean-convex boundary.
(a)
If is stable, then it is a disc and .
(b)
If has index , then .
Remark 1.1.
The scalar curvature in [6] is one half of ours. This is why a factor appears in item (iii) of Theorem E instead of the of Theorem 1.2 in [6].
Acknowledgements
The author would like to thank Paolo Piccione for his constant support and encouragement during the period when this article was written and revised, and for countless fruitful mathematical conversations. He also expresses sincere gratitude to Lucas Ambrozio for valuable comments on this work. Additionally, the author thanks Izabella Freitas and Jackeline Conrado for the figures in this paper.
2. Preliminaries: variational problem and stability
The purpose of this section is to formally introduce the concept of capillary CMC and minimal hypersurfaces.
Despite the fact that in section 3 we will be dealing only with capillary surfaces in Riemannian -manifolds, there is no significant simplification in introducing the main concepts only in dimension . So, the general situation will be addresed in the sequel.
Let be a Riemannian manifold with nonempty boundary. Let be a smooth compact manifold with nonempty boundary, and let be a smooth immersion of into . We say that is a proper immersion if .
Henceforth, we assume that is two-sided. Fix a unit normal vector field for along and denote by the outward unit conormal for in . Moreover, let the outward pointing unit normal for and let the unit normal for in such that the bases and determine the same orientation in . See Figure 2 to gain some intuition.
A smooth function is called a proper variation of is the maps , defined by , are proper immersions for all , and if .
Let us fix a proper variation of . The variational vector field associated to is the vector field along defined by
We now define some important functionals related to the variation . The area functional is given by
where denotes the area element of . Even if , it is customary to refer to this as the area functional.
The volume functional is defined by
where is the volume element of . We say that the variation is volume preserving if for every .
We also consider the wetting area functional :
where denotes the area element of .
Finally, we define the energy functional. In order to do so, let us fix an angle . Then is given by
The following proposition contains the formulae for the first variation of the energy and volume, whose proof can be found in [20] and [5]:
Proposition 2.1.
Let be a proper variation of the immersion . Then the following formulae hold:
where , is the mean curvature of with respect to , is the area element of induced by and is the line element of induced by .
We say that the immersion is a capillary CMC immersion if for every volume preserving variation of . If for every variation of , we call a capillary minimal immersion. When there is only one immersion under consideration and there is no risk of confusion, we just say that is a capillary CMC or minimal hypersurface. See Figure 2 for an example in an Euclidean cone.
Figure 2. The four fundamental vector fields in a capillary minimal surface in a cone.
Notice that is a capillary CMC hypersurface if and only if has constant mean curvature and along ; this last condition means that meets at an angle of . Similarly, is a capillary minimal hypersurface when is a minimal hypersurface and meets at an angle of . When , we use the term free boundary CMC (or minimal) hypersurface.
For a capillary CMC or minimal hypersuface with contact angle , the orthonormal bases and are related by the following equations:
This will be important in the calculations.
Proposition 2.2.
[1, Proposition 2.1]
Let be a proper immersion which is transversal to . For any smooth function satisfying , there exists a volume preserving variation of such that , where is the variational vector field associated to . If we don’t assume that , then the result still holds, but now doesn’t need to be volume preserving.
We now discuss the notion of stability. Firstly, let us fix some notation. For a proper immersion , let the second fundamental form of , and denote by the Laplacian of with respect to the metric induced by . Moreover, let be the second fundamental form of with respect to .
The following proposition gives a formula for the second variation of energy. The proof is long and can be found in the Appendix of [18].
Proposition 2.3.
Let be a capillary CMC immersion and let be a smooth function which satisfies . Let be a volume preserving proper variation of such that (see Proposition 2.2). Then,
where
and is the Jacobi operator of . If is a capillary minimal immersion and is any smooth function defined on , then the same formula holds, but now doesn’t need to be volume preserving.
The capillary CMC immersion (or just ) is called stable if for any volume preserving variation of . If is a capillary minimal immersion, we call it stable whenever for every variation of .
Alternatively, let , where is the first Sobolev space of . The index form of is given by
(1)
Then is a capillary CMC stable immersion if and only if for every . If is a capillary minimal immersion, then it is stable precisely when for every .
The stability index of a capillary CMC (resp. minimal) hypersurface is the dimension of the largest vector subspace of (resp. ) restricted to which the bilinear form is negative definite. Thus, stable hyperfurfaces are those which have index equal to zero. Intuitively, this means that no variation of a capillary minimal stable hypersurface decreases area up to second order.
It is possible to analitically compute the stability index of a capillary CMC (resp. minimal) hypersurface in terms of the spectrum of its Jacobi operator with a Robin boundary condition. Indeed, the boundary condition
is elliptic for the Jacobi operator . Therefore, there exists a nondecreasing and divergent sequence of eigenvalues for the problem
(2)
The index of is then the number of negative eigenvalues of the problem above associated to (smooth) eigenfunctions that lie in (resp. ) (see more details in [6] and [19]).
3. Proof of the Theorems
This section is devoted to prove the Theorems mentioned in the Introduction.
Recall that, given a Riemannian -manifold with nonempty boundary and a fixed angle , we can define the following functional in the space of compact properly immeresd surfaces in :
provided that the scalar curvature and the mean curvature are bounded from below.
We first show that if is capillary minimal stable, then can be bounded above by the topology of . We also characterise the equality case.
We begin with a simple lemma.
Lemma 3.1.
Let be a Riemannian -manifold with nonempty boundary, and let be a two-sided capillary CMC surface embedded in with contact angle and mean curvature equal to . Then:
Proof.
Let be a unit tangent vector field along . Since is CMC, we know that . So,
But is an orthonormal referential for along . Thus,
Let us assume now that equality holds. It is immediate to see that must be totally geodesic, item (b) holds and . The latter implies that for any . Indeed, if , then
Dividing by and rearranging, we obtain
Now we let and conclude that . The reverse inequality is analogous. This proves the claim. Thus, we know that
for any . This implies that and . Since we already know that , it follows that , proving item (c).
It remains to calculate the geodesic curvature of in . We have
But Lemma 3.1 implies that . So, , proving (a). The converse statement can be easily proved.
∎
We are amost ready to prove Theorem B. Recall that is a vector field in which is tangent to along the entire boundary of and which satisfies along . Let be the local flow of and fix .
We begin with a lemma, whose proof can be found in the appendices of [2] and [12].
Lemma 3.2.
Let be a Riemannian -manifold with nonempty boundary, and let a two-sided proper embedding of a compact surface with boundary. Fix a proper variation of and let . We use the subscript to denote all terms related to , e.g. is the unit normal, is the mean curvature, etc. Let and let be the contact angle between and . Then the following formulae hold:
where denotes the orthogonal projection onto , is the Jacobi operator of and is a certain vector field along . In particular, if each is capillary CMC with contact angle , then
Given a function , let . If the norm of is small enough, then is a properly embedded surface in . As in the statement of Lemma 3.2 we use the subscript to denote the terms related to .
Now consider the Banach spaces
For small and , define the map by
where denotes the open ball of radius centered at the origin of . We claim that is an isomorphism when restricted to . To prove this, let . Notice that the map is a proper variation of whose variational vector field is
So, using the formulae from Lemma 3.2 and the fact that the Jacobi operator of is just the Laplacian, since is infinitesimally rigid, we obtain:
(6)
(7)
If this is equal to zero, then is harmonic and satisfies a Neumann condition on the boundary of . This implies that is constant, and this constant must be equal to zero since , which proves that is injective. The surjectivity follows from classical results for Neumann type boundary conditions for the Laplace operator.
The final step is to apply the implicit function theorem: for some smaller , there exists a function such that and for any . This means precisely that the surfaces
have constant mean curvature and meet along their boundaries at an angle of .
Now let be given by . Then and . Define by and observe that this is a proper variation of whose variational vector field is
Since
for all , differentiating this equation with respect to implies that satisfies the Neumann problem on (see equation (6)). Therefore, it must be constant. But
since . By differentiating this equation with respect to and evaluating it at we obtain that . Thus, as we wanted to show.
Finally, since for every , is transverse to and is properly embedded, by taking a smaller if necessary, we may assume that parametrises a foliation of a neighbourhood of in . This concludes the proof.
∎
The proof of Theorem C is similar to the proof of Theorem B. This way, we only sketch it, indicating the main differences.
where denotes the open ball of radius centered at the origin of . Then, using the formulae from Lemma 3.2 and the fact that the Jacobi operator of is just the Laplacian, since is infinitesimally rigid, we have:
(8)
for any . It is easy to show that is an isomorphism when restricted to . Now, just apply the implicit function theorem and proceed as in the proof of Theorem B.
∎
Before proving Theorem D, let us clarify that a closed curve in is locally length-minimising when every nearby closed curve in has length greater than or equal to the length of .
Let us start with a lemma.
Lemma 3.3.
Let be a Riemannian -manifold with nonempty boundary and let be a two-sided capillary minimal proper immersion of a compact surface , with contact angle equal to . If is energy-minimising for the angle , then, changing the sign of its unit normal, it is energy-minimising for the angle .
Proof.
Given a proper variation of , define to be the variation given by . Denote by and the wetting areas of and , respectively. We claim the for every . To prove this, define by
where is a unit tangent vector field along and is the standard vector field in . Then, for , we have
The computation is analogous for . This proves the claim. To prove the lemma, we show that is equal to :
Thus, we have
for any variation and any , which implies that is energy-minimising for the angle .
∎
Let be a parametrisation of the capillary CMC foliation given by Theorem B. For each , let , where as usual, we use the subscript for terms related to . Then, by equations (4) and (5) in Lemma 3.2, we have
(9)
(10)
Notice that , so we can assume for any , maybe for some smaller . Using the Gauss equation, we rewrite equation (9) as
where denotes the scalar curvature of restricted to and denotes the Gaussian curvature of . Now recall that is constant on , so, integration by parts gives
We now apply Gauss-Bonnet Theorem and Lemma 3.1 to obtain
Since each is homeomorphic to , . Next, we use the fact that is infinitesimally rigid to apply the identity :
By hypothesis, . If each component of is length-minimising, then the second term in the right hand side is nonnegative, and if , then it is equal to zero. In either case, we have
By the first variation of area,
(11)
Thus, we obtain the following differential inequality:
Let us rewrite this as
(12)
where
•
•
•
•
It is time to apply Lemma 3.3. Changing the sign of the unit normal of in the begining of the proof, if necessary, we may assume that if and that if . In either case, inequality (12) implies that . If then for all small positive . The first variation formula for the energy (see Proposition 2.1) would then give that :
which is absurd, since is energy-minimising. So . Now let denote the real function in the right hand side of inequality (12). A simple computation yields . This is positive if and only if and . In this case, we would have , which is also absurd by the same reason. So either , and we are done, or .
Let us analyse the case . Inequality (12) takes the form , and since , this implies that for and for . By the first variation of energy again, for any . Thus, because is energy-minimising. So, by the same formula. In particular, every is capillary minimal stable. Applying Theorem A we obtain that, for every ,
This way, every inequality above is an equality, which shows that every is infinitesimally rigid. In particular, has constant geodesic curvature equal to in . Since we are supposing either condition (a) or (b) in the theorem, we conclude that either or each component of is a geodesic in . Hence, either and we are done, or .
Assuming and we are going to show that is flat in a neighbourhood of and that is totally geodesic in this neighbourhood.
Claim: The unit normals to define a parallel vector field.
Firstly, notice that each map is constant because it harmonic with Neumann boundary condition by equations (9) and (10). Now, let be local coordinates on and write . Since each is totally geodesic, we know that . Moreover, since is of unit length, is tangent to . Thus,
This proves the Claim.
The Gauss equation for shows that , where and denote the sectional curvatures of and of . This holds since is flat and totally geodesic. Now, since is parallel, . The symmetries of the curvature tensor then imply that , that is, is flat around .
Finally, let us show that is totally geodesic near . By infinitesimal rigidity we know that for every . If we denote by a unit normal for , it is straighforward to check that and . Since is an orthonormal referential for around , the bilinearity of implies that , i.e. is totally geodesic in the union of all , .
The last statement of the theorem (when ) follows from Theorem 7 in [2].
∎
Remark 3.4.
Theorem D can be proved using the capillary minimal foliation with variable contact angle given by Theorem C instead of the capillary CMC foliation of Theorem B.
Finally, we move on to Theorem E, which gives conditions under which we can control the topology of low index surfaces in some Riemannian -manifolds.
As usual, we start with a lemma, whose proof can be found in [6].
Lemma 3.5.
Let be a closed orientable Riemannian surface of genus , and let be any nonnegative smooth function. Then there exists a conformal map such that and has degree less than or equal to , where denotes the integer part of .
Let a local orthonormal frame on such that and are tangent to and coincides with the unit normal to . Since is minimal, the Gauss equation gives that
So,
(13)
Now let be the first eighenfunction of Problem (2). By gluing a disc on each boundary component of , we may view as a compact domain of a closed orientable Riemannian surface . So, Lemma 3.5 furnishes a conformal map such that and . Let denote the restriction of to . If we write , then we know that each component is orthogonal to the first eigenfunction . Since has index , we have:
Summing over and since , we get
By the conformality of ,
(14)
Therefore,
We now use equation (13) and the hypotheses of curvature and to continue:
as we wanted. The particular case mentioned in the theorem follows immediately.
(ii)
To prove this item, we use the identity
(15)
Let be the same function of item (i). As before,
Using equation (15) and inequality (14), we obtain
Now we use Gauss-Bonnet therorem and Lemma 3.1 to write
Finally, we use the curvature assumptions to get
as we wanted.
(iii)
Item (a) follows from Theorem A, since under the curvature hypotheses must be a disc. Item (b) follows from from item (ii) of the current theorem (again, must be a disc).
∎
References
[1]
A. Ainouz and R. Souam,
Stable Capillary Hypersurfaces in a Half-Space or a Slab,
Indiana Univ. Math. J. 65(3)(2016), 813–831.
[2]
L. Ambrozio,
Rigidity of area-minimizing free boundary surfaces in mean convex three-manifolds,
J. Geom. Anal. 25 (2015), 1001–1017.
[3]
L. Ambrozio, A. Carlotto and B. Sharp,
Compactness analysis for free boundary minimal hypersurfaces,
Calc. Var. PDE 57(2018), 1–39.
[4]
L. Ambrozio, A. Carlotto and B. Sharp,
Index estimates for free boundary minimal hypersurfaces,
Math. Ann. 370(3–4)(2018), 1063–1078.
[5]
J. L. Barbosa, M. do Carmo and J. Eschenburg,
Stability of hypersurfaces of constant mean curvature in Riemannian manifolds,
Math. Z. 197(1988), 123–128.
[6]
J. Chen, A. Fraser and C. Pang,
Minimal immersions of compact bordered Riemann surfaces with free boundary,
Trans. Amer. Math. Soc. 367(4)(2014), 2487–2507.
[7]
J. Choe and M. Koiso,
Stable capillary hypersurfaces in a wedge,
Pacific J. Math. 280(1)(2016), 1–15.
[8]
R. Courant,
The Existence of a Minimal Surface of Least Area Bounded by Prescribed Jordan Arcs and Prescribed Surfaces,
Proc. Natl. Acad. Sci. USA 24(2)(1938), 97–101.
[9]
R. Courant,
Dirichlet’s Principle, Conformal Mapping, and Minimal Surfaces,
Springer-Verlag New York, 1950.
[10]
R. Finn,
Equilibrium Capillary Surfaces,
Volume 284 of Grundlehren der mathematischen Wissenschaften, Springer-Verlag New York, 1986.
[11]
J. L. Lagrange,
Essai d’une nouvelle méthode pour déterminer les maxima et les minima des formules intégrales indéfinies,
Miscellanea Taurinensia 2(1)(1760), 173–199.
[12]
C. Li,
A polyhedron comparison theorem for -manifolds with positive scalar curvature,
Invent. Math. 219(2020), 1–37.
[13]
R. López,
Stability and bifurcation of a capillary surface on a cylinder,
SIAM J. Appl. Math. 77(1)(2017), 108–127.
[14]
J. B. Meusnier,
Mémoire sur la courbure des surfaces,
Mém. Mathém. Phys. Acad. Sci. Paris, prés. par div. Savans 10(1776), Presented in 1776, 477–510.
[15]
J. C. C. Nitsche,
Stationary partitioning of convex bodies,
Arch. Rational Mech. Anal. 89 (1985), 1–19.
[16]
S. H. Park,
Every ring type spanner in a wedge is spherical,
Math. Ann. 332(3)(2005), 475–482.
[17]
M. Ritoré and C. Rosales,
Existence and characterization of regions minimizing perimeter under a volume constraint inside
Euclidean cones,
Trans. Am. Math. Soc. 356(11)(2004), 4601–4622.
[18]
A. Ros and R. Souam,
On stability of capillary surfaces in a ball,
Pacific J. Math. 178(2) (1997), 345–361.
[19]
R. Schoen,
Minimal submanifolds in higher codimension,
Mat. Contemp. 30(2006), 169–199.
[20]
M. Spivak,
A Comprehensive Introduction to Differential Geometry,
Volume 4, Publish Or Perish, 1999.
[21]
G. Wang and C. Xia,
Uniqueness of stable capillary hypersurfaces in a ball,
Math. Ann. 374 (2019), 1845–1882.