Heintze-Karcher inequality and
capillary hypersurfaces in a wedge
Abstract.
In this paper, we utilize the method of Heintze-Karcher to prove a "best" version of Heintze-Karcher-type inequality for capillary hypersurfaces in the half-space or in a wedge. One of new crucial ingredients in the proof is modified parallel hypersurfaces which are very natural to be used to study capillary hypersurfaces. A more technical part is a subtle analysis along the edge of a wedge. As an application, we classify completely embedded capillary constant mean curvature hypersurfaces that hit the edge in a wedge, which is a subtler case.
MSC 2020: 53C24, 35J25, 53C21
Keywords: Heintze-Karcher’s inequality,
capillary hypersurface, CMC hypersurface, Alexandrov’s theorem.
1. Introduction
The study of capillary surfaces goes back to Thomas Young, who studied in 1805 the equilibrium state of liquid fluids. It was he who first introduced the notion of mean curvature and the boundary contact angle condition of capillarity, the so-called Young’s law. This problem was reintroduced and reformulated by Laplace and by Gauss later. For the history of capillary surfaces see Finn’s survey [Finn99]. A capillary hypersurface in with boundary on a support hypersurface is a critical point of the following functional
among all compact hypersurfaces with boundary on under a volume constraint. Here is the bounded domain enclosed by and , and is the -dimensional area of , and . Equivalently, a capillary hypersurface is a constant mean curvature (CMC) hypersurface with boundary which intersects the support hypersurface at a constant angle . There has been a lot of interdisciplinary investigations on the stationary solutions and local minimizers of the above energy. For the interested reader, we refer to Finn’s book [Finn86], which is an excellent survey on capillary surfaces.
Inspired by the recent development of the min-max theory for minimal surfaces and CMC surfaces [CD03, MN16a, MN16b], there have been a lot of works on free boundary minimal surfaces and CMC surfaces, which are a special class of capillary surfaces with , see for example [DR18, LZ21, GLWZ21, ZZ19]. Very recently the min-max theory for capillary surfaces was developed in [DMDP21, LiZZ21].
One important application of the capillary surfaces was recently obtained by Chao Li in [Li21], where he utilizes capillary surfaces in a polyhedron to study Gromov’s dihedral rigidity conjecture. His work on capillary surfaces is related to this paper, in which we will consider capillary surfaces supported on a wedge.
The main objective of this paper is to make a complete classification of embedded capillary hypersurfaces in a wedge. Such a hypersurface can be seen as a model for capillary hypersurfaces in Riemannian polyhedra. For the precise definition of a wedge, see below in the Introduction. Our starting point is the Heintze-Karcher inequality. Let us first recall Heintze-Karcher’s theorem and Heintze-Karcher’s inequality for closed hypersurfaces.
In a seminal paper [HK78], Heintze-Karcher proved a general tubular volume comparison theorem for embedded Riemannian submanifolds, which generalizes the celebrated Bishop-Gromov’s volume comparison theorem in Riemannian geometry. For an embedded closed hypersurface , which encloses a bounded domain , in an -dimensional Riemannian manifold of nonnegative Ricci curvature, Heintze-Karcher’s theorem reads as follows,
(1) |
Here is the mean curvature of at and is the length to reach the first focal point of from by the normal exponential map. As a direct consequence of (1), one deduces that
(2) |
provided that is strictly mean convex, namely, on . Nowadays, (2) is literately referred to as Heintze-Karcher’s inequality in hypersurfaces theory. A well-known new proof via Reilly’s formula [Reilly77] has been given by Ros [Ros87]. Moreover, Ros [Ros87] utilized the Heintze-Karcher inequality (2) to reprove the celebrated Alexandrov’s soap bubble theorem, which states that any embedded closed constant mean curvature hypersurfaces in must be a round sphere.
Since then various Heintze-Karcher-type inequalities have been established in various circumstance. For instance, Montiel-Ros [MR91] and Brendle [Br13] established Heintze-Karcher-type inequalities in space forms and in certain warped product manifolds respectively, see also [QX15, LX19]. The Heintze-Karcher inequality in has been also established for sets of finite perimeter, see e.g. [DM19, San19]. Like the Alexandrov-Fenchel inequalities, the Heintze-Karcher inequality becomes one of fundamental geometric inequalities in differential geometry.
Inspired by the method of Ros [Ros87], and also by the work of Brendle [Br13], we have proved a Heintze-Karcher-type inequality for hypersurfaces with free boundary in a unit ball [WX19] by using a generalized Reilly’s formula proved by Qiu-Xia [QX15]. However this method leads to a slight different inequality if we consider hypersurfaces with capillary boundary in the unit ball or in the half-space in the very recent work, [JXZ22]. Precisely we have established in the previous work [JXZ22] a version of Heintze-Karcher-type inequality for hypersurfaces in the half-space with capillary boundary, by using the solution to a mixed boundary value problem in the classical Reilly’s formula. Let , where , and for an embedded, compact, strictly mean-convex hypersurface with capillary boundary with a constant contact angle , there holds
(3) |
with equality if and only if is a spherical cap.
Inequality (3) is optimal, in the sense that the spherical caps achieve equality in (3). However it is not in the best form. For example, while we are able to use (3) to reprove the Alexandrov theorem for constant mean curvature (CMC) hypersurfaces in [JXZ22], it is not very helpful to handle the case of higher order mean curvatures (see (17) for the definition). In view of the following Minkowski formula
(4) |
a possible best form, which was conjectured in [JXZ22], is
(5) |
It is clear by using the Cauchy-Schwarz inequality that inequality (5) implies (3), provided that is non-negative. But without the non-negativity one does not know which one is stronger.
The first part of this paper is to establish this “best” version of the Heintze-Karcher inequality in a little more general setting and for whole range . Let be a hypersurface in with (possibly non-connected) boundary . The hypersurface intersects with the supported hyperplane transversely. 111In the paper we abuse a little bit the terminology of capillarity. A capillary hypersurface mentioned at the beginning of the Introduction is called a capillary CMC hypersurface in the paper.
Theorem 1.1.
Let and let be a smooth, compact, embedded, strictly mean convex -capillary hypersurface, with for every . Let denote the enclosed domain by and . Then it holds
(6) |
Equality in (6) holds if and only if is a -capillary spherical cap.
Note that we also have removed the restriction that , comparing with the previous work [JXZ22]. As a direct application, we get the Alexandrov-type theorem for embedded capillary hypersurfaces in with constant -th mean curvature, for any .
Corollary 1.2.
Let and . Let be a smooth, embedded, compact, -capillary hypersurface with constant -th mean curvature. Then is a -capillary spherical cap.
Our proof of Theorem 1.1 is inspired by the original idea of Heintze-Karcher [HK78] (see also Montiel-Ros [MR91]) which uses parallel hypersurfaces to estimate the enclosed volume. However the ordinary parallel hypersurfaces do not work for capillary hypersurfaces. The one of key ingredients of this paper is a correct form of parallel hypersurfaces defined in (33). To prove Theorem 1.1, we need to show the surjectivity of onto the enclosed domain , for which we discover an appropriate foliation by round spheres with simultaneously varied center and radius.
It is interesting to see that our proof of Theorem 1.1 provides a refinement of the ordinary Heintze-Karcher inequality for closed hypersurfaces, since any closed hypersurface can be viewed as a capillary hypersurface with an empty boundary. Hence we have
Corollary 1.3.
Let be a closed, strictly mean convex hypersurface in with enclosed domain . Then it holds
Equality in (1.1) holds if and only if is a round sphere.
In the second part of this paper, we study hypersurfaces with capillary boundary in a wedge domain. Here we simply call it a wedge. An ordinary wedge, we call it a classical wedge in this paper, is the unbounded closed region determined by two intersecting hyperplanes with dihedral angle , which is also called an opening angle, lying in . There have been many works on the study of the stability of CMC capillary hypersurfaces (c.f. [CK16, LX17, Souam21, XZ21]) and on embedded CMC capillary hypersurfaces in wedges (c.f. [McCuan97, Park05, Lopez14]). Comparing with the half-space case, a big difference is that the Alexandrov’s reflection method might fail in the case of wedges, though the authors in [McCuan97, Park05, Lopez14] managed to modify Alexandrov’s reflection to obtain their classification results in certain cases. It is interesting that our method to establish the Heintze-Karcher-type inequality works in the wedge case, and even works in a more general setting. See also the recent development of this method in the anisotropic setting [JWXZ23, JWXZ23b].
In fact we shall consider generalized wedges which are determined by finite many mutually intersecting hyperplanes. To be more precise, let be the unbounded closed region in , which are determined by finite many mutually intersecting hyperplanes , for some integer , such that the dihedral angle between and , , lies in . We call such a generalized wedge. is called an edge of the wedge . If , we call a classical wedge. Let be the outwards pointing unit normal to in for . Thus are linearly independent. Up to a translation, we may assume that the origin . Given . Now we define an important vector associated with and by
(7) |
where is such that . We say that is a -capillary hypersurface in with if it intersects at contact angle for . It is easy to see that is the center of the -capillary spherical cap with radius . The following key assumption (8) has a clear geometric meaning that the unit sphere centered at intersects the edge of the wedge .
We shall prove the following Heintze-Karcher-type inequality in a wedge.
Theorem 1.4.
Let be a generalized wedge whose boundary consists of mutually intersecting closed hyperplanes and . Assume that
(8) |
Let be a smooth, compact, embedded, strictly mean convex -capillary hypersurface with for , . Let be the enclosed domain by and . Assume in addition that does not hit the edges of , i.e.,
(9) |
Then
(10) |
with equality if and only if is a -capillary spherical cap.
The idea of proof of Theorem 1.4 is similar to that of Theorem 1.1, by using a suitable family of parallel hypersurfaces which relates . To show the surjectivity of , assumption (9) plays a crucial role. Actually, this condition was required in previous related papers, except [Lopez14]. See Remark 1.10. In this paper we are able to remove the additional assumption (9) in a classical wedge, i.e., . Precisely, we have the following
Theorem 1.5.
When , Theorem 1.4 holds true without assumption (9).
The proof of Theorem 1.5 relies on a delicate analysis on the edge, which is the most technical part of this paper. A special case, and , i.e., is a free boundary hypersurface, for which (10) was proved by Lopez in [Lopez14] via Reilly’s formula. It is a natural question to ask if Theorem 1.4 holds true for without (9). Theorem 1.5 leads us to believe that assumption (9) is unnecessary.
Now we make some remarks on condition (8).
Remark 1.6.
As applications of Theorem 1.4 and Theorem 1.5, we prove an Alexandrov-type theorem and a non-existence result for embedded CMC capillary hypersurfaces in a wedge.
Theorem 1.7.
Let be a classical wedge whose boundary consists of with . Let be a smooth, compact and embedded -capillary hypersurface with constant -mean curvature, . Assume . Then is a -capillary spherical cap.
Theorem 1.8.
Let be a classical wedge whose boundary consists of with . Then there exists no smooth, compact and embedded -capillary, CMC hypersurface such that and . Moreover, there exists no smooth, compact, embedded, -capillary hypersurface of constant -mean curvature for some , such that and .
As a consequence of Theorem 1.7, Theorem 1.8 and Remark 1.6 (iii), we have the following
Theorem 1.9.
Let be a classical wedge whose boundary consists of with . Let be a smooth, compact and embedded -capillary CMC hypersurface. Then is a -capillary spherical cap which intersects with the edge if and only if .
Several remarks and questions are in order.
Remark 1.10.
-
(i)
McCuan [McCuan97, Theorem 2] proved that any -capillary spherical cap which is disjoint with the edge must satisfy . This condition implies that , see Lemma A.1.
-
(ii)
Lopez [Lopez14] proved an Alexandrov-type theorem for embedded CMC capillary surfaces with , i.e., the free boundary case. Note that in this case, is automatically satisfied. Hence Theorem 1.9 covers Lopez’s result.
In contrast to it a ring-type CMC free hypersurface in a wedge was constructed by Wente in [Wente95], which is certainly not embedded. It is natural to ask whether there exist immersed ring-type CMC -hypersurfaces for a general .
-
(iii)
Park [Park05] classified the embedded CMC capillary ring-type spanners, which are topologically annuli and disjoint with the edge. Our Theorem 1.7 classified all embedded CMC capillary surfaces intersecting with the edge, without any topological condition.
-
(iv)
McCuan [McCuan97] proved a non-existence result for the embedded CMC capillary ring-type spanners with
(12) by developing spherical reflection technique, when . Note that the angle relation (12) is weaker than , see Lemma A.1. However, our Theorem 1.8 requires no topological assumption. Moreover it holds for any dimensions.
-
(v)
For stable CMC capillary hypersurfaces in a classical wedge, Choe-Koiso [CK16] proved that such a surface is a part of a sphere without the angle condition (11), but with condition (9) and with the embeddness of for or the convexity of for . It is an interesting question to ask if an immersed stable CMC capillary hypersurface in a wedge is a part of a sphere, without any further conditions, c.f., [WX19, GWX21, Souam21].
We end the Introduction with a few supplement on the study of capillary hypersurfaces in a wedge domain. A nonparametric capillary surface is a graph of a function over a domain, say , which satisfies the constant mean curvature equation with a corresponding capillary boundary condition. This is an equilibrium free surface of a fluid in a cylindrical container. When the domain has a corner, then this nonparametric surface can be viewed as a capillary surface in a wedge (or in a wedge domain). There have been a lot of research on such a problem, especially after Concus-Finn [CF69], where it was already observed that the opening angle of the wedge and both contact angles should satisfy certain conditions for the existence. See also [CF74]. Later in [CF96] Concus-Finn proved that is continuous at a given corner if (11) holds, while if (11) does not hold there is no solution in one case and in the left case, namely , they conjectured that has a jump discontinuity at the corner. See the Concus-Finn rectangle in [Lan10], Figure 2. This conjecture was solved by Lancaster in [Lan10] with the methods developed by Allard [A72] and especially by Simon [S80]. The latter was crucially used in a very recent work of Chao Li [Li21] mentioned at the beginning of the Introduction. See also his further work [EL22] with Edelen on surfaces in a polyhedral domain, which is also closely related to surfaces in a wedge domain.
The rest of the paper is organized as follows. In Section 2, we collect some basic facts about wedges and capillary hypersurfaces in wedges. In Section 3 and Section 4, we prove the main theorems on the Heintze-Karcher inequality in the half-space and a wedge. In Section 5, we prove the Alexandrov-type theorem and the non-existence result for CMC capillary hypersurfaces in a wedge.
2. Notations and Preliminaries
Let be the bounded domain in with piecewise smooth boundary , where is a smooth compact embedded -capillary hypersurface in and . Denote the corners by , which are smooth, co-dimension two submanifolds in . For the sake of simplicity, we denote by the union of , i.e., . We use the following notation for normal vector fields. Let and be the outward unit normal to and (with respect to ) respectively. Let be the outward unit co-normal to and be the outward unit co-normal to . Under this convention, along each and span the same 2-dimensional plane and have the same orientation in the normal bundle of . Hence one can define the contact angle function along each , by
(13) | |||
(14) |
Let denote the -tuple . We call a -capillary hypersurface, if we want to emphasize the contact angle function. We also use -capillary hypersurface to denote such a hypersurface with , a vector
We denote by , , and , the gradient, the Laplacian, the Hessian and the divergence on respectively, while by , , and , the gradient, the Laplacian, the Hessian and the divergence on the smooth part of , respectively. Let , and be the first, second fundamental forms and the mean curvature of the smooth part of respectively. Precisely, and . In particular, since is planar, the second fundamental form , correspondingly, the mean curvature vanishes.
We need the following structural lemma for compact hypersurfaces in with boundary, which is well-known and widely used, see [AS16, JXZ22].
Lemma 2.1.
Let be a smooth compact hypersurface with boundary. Then it holds that
(15) |
Proof.
Let for any constant vector , where and denote the tangential component of and respectively. One computes that
Integration by parts yields the assertion. ∎
The following lemma is well-known when the capillary hypersurfaces are bounded by containers with totally umbilical boundaries, in particular, a wedge in , see e.g., [AS16, LX17, WX19].
Lemma 2.2.
Let be a wedge and for . If is a smooth -capillary hypersurface, then along , is a principal direction of .
Proof.
For the completeness we provide a proof. It suffices to prove that for any vector tangent to . Indeed,
(16) |
where we have used (13), the constancy of , the fact that are unit vector fields, and since are totally geodesic. This completes the proof. ∎
The -th mean curvature of is defined by the identity:
(17) |
for all real number . Thus is the mean curvature of and is the Gaussian curvature, and we adopt the convention that .
We have the following Minkowski-type formula for -capillary hypersurfaces.
Proposition 2.3.
Let be a wedge and be a -capillary hypersurface. Then it holds that for ,
(18) |
In particular, if , i.e. , then
(19) |
Proof.
The case for has been proved in [LX17, Lemma 5]. For the sake of completeness, we include the proof here.
Since
by using integration by parts in , we get
(20) |
From the capillary boundary condition (13) it is easy to see that on each
(21) |
By (15), (13) and (21), we get
(22) |
It follows from (20) and (2) that
(23) |
Now we prove (18) for general . For a small real number , consider a family of hypersurfaces with boundary , defined by
We claim that is also a -capillary hypersurface in . In fact, if are principal directions of a point of and are the corresponding principal curvatures, we have
(24) |
From (24), we see that , where denotes the outward unit normal of at . Moreover, the capillarity condition (14) implies: for any , we have
(25) |
in other words, , and hence . In view of this, we have: ; that is, is also a -capillary hypersurface in .
Therefore, we can exploit (23) to find that
(26) |
By (24), the tangential Jacobian of along at is just
(27) |
where is the polynomial defined in (17). Moreover, using (24) again, we see that the corresponding principal curvatures are given by
(28) |
Hence fix , the mean curvature of at , say , is given by
(29) |
where is the -th mean curvature of at .
We have also
(32) |
which is easy to prove. If one views (32) as one of (18) with , the Heintze-Karcher inequality (6) that we want to prove could also be viewed one of them with . Certainly now it is an inequality, instead of an equality.
Remark 2.4.
An alternative proof of Proposition 2.3 can be given as that of [WWX22, Proposition 2.5], where the Minkowski-type formula for the half-space case has been proved.
In the sequel, will be always referred to as a smooth, embedded capillary hypersurface.
3. Heintze-Karcher Inequality in the Half-Space
Proof of Theorem 1.1.
Let be a -capillary hypersurface with along . For any , let be the set of unit principal vectors of at and the set of corresponding principal curvatures. Since is strictly mean convex,
We define
and
(33) |
gives a family of hypersurfaces , which are the modified parallel hypersurfaces mentioned above.
Claim: .
Indeed, let us denote by the closed ball centered at of radius , and . For any , we consider a family of spheres . Since is an interior point, when is small enough, we have . Since , it is easy to see that the spheres gives a foliation of . Hence must touch as we increase the radius . As a conclusion, for any , there exists and , such that touches for the first time, at the point . We have two cases.
Case 1. .
In this case, since , the sphere is tangent to at from the interior. It follows that . Invoking the definition of and , we find that in this case.
Case 2. .
We will rule out this case by the condition on the contact angle function of . In this case, by the first touching property of , the contact angle of with is smaller than or equals to , which is smaller than or equals to , by assumption. (see Figure 1 for an illustration). However implies that , a contradiction to . The Claim is thus proved.

By a simple computation, we find
Hence the tangential Jacobian of along , at is just
By virtue of the fact that , the area formula yields
By the AM-GM inequality, on , and the fact that , we obtain
which is (6).
The characterization of equality case in (6) follows from the classical one. Precisely, since the equality holds throughout the argument, the arithmetic mean-geometric mean (AM-GM) inequality assures the umbilicity of , and it follows that is a spherical cap. Apparently, the contact angle of a spherical cap with a hyperplane is a constant, say . It is easy to see that , and hence must be a -capillary spherical cap. Conversely, when is a -capillary spherical cap, then is a positive constant. By virtue of the Minkowski formula (19) for , we see that equality in (6) holds. ∎
Let us close this section with a remark. In the proof of , our choice of the touching balls is enlightened by the following observation.
Remark 3.1 (Foliation of -capillary hypersurfaces).
In [MR91], to prove the Heintze-Karcher inequality for closed hypersurfaces, one shall ‘sweepout’ the domain by a foliation around any point , whose leaves are level-sets of the distance function to . The key point is that, such ‘sweep-outs’ coincides with the domain , if and only if is a ball and is chosen to be the center.
In view of this, our choice of foliation in the capillary case is thus clear; we want to ‘sweepout’ the -ball with the foliation, whose leaves are -spherical caps(as illustrated in Figure 2).

(a) Sweepout of -domain.

(b) Sweepout of generic domain.
4. Heintze-Karcher Inequality in a Wedge
In this section, we prove Theorem 1.4 and Theorem 1.5 in a wedge , following largely from the proof of the Heintze-Karcher-type inequality presented in the previous section.
Proof of Theorem 1.4.
Let be a compact embedded hypersurface with -capillary boundary, where for each and every . As above, for any , let be the set of principal unit vectors of at and the set of the corresponding principal curvatures. Now we define modified parallel hypersurfaces by
As in Theorem 1.1, we shall show that . Let . We consider the sphere foliation . By virtue of Lemma A.3, there exists some such that touches from the interior at a first touching point .
Case 1. We can get , as argued in Case 1, Theorem 1.1.
Case 2. . By assumption (9), for some . We will rule out this case again by virtue of the capillarity of . In this case, is tangent to , and the touching angle of with at must be smaller than , and it follows from the geometric relation that lies outside . Precisely, up to a rotation, we may assume that the touching plane is , say , and we denote by the touching angle, satisfying , due to the first touch. From the geometric relation(see Figure 3), we find, , the angle relation then implies that 222Notice that , which means moving along with distance is indeed moving along with distance ., which contradicts to the fact that . Therefore, we complete the proof that .

By a simple computation as Theorem 1.1, we see, the tangential Jacobian of along at is just
By a similar argument as Theorem 1.1, we conclude
(34) |
As proved in Theorem 1.1, if equality in (10) holds, then is umbilical, and hence spherical. To see that must be a -capillary spherical cap, we need a different argument. As equalities hold throughout the argument, we have
(35) |
Moreover, (14) implies: for any , there holds
(36) |
Recall that is the outwards pointing unit normal of , and we have already showed in the previous proof that . Thus, if strictly at some , then it must be that , which contradicts to (35). In other words, for any , we must have , this shows that must be a -capillary spherical cap. ∎
Proof of Theorem 1.5.
We note that the proof follows closely the one of Theorem 1.4. Precisely, thanks to Lemma A.3, we can use our foliation to test the surjectivity of , i.e., . One subtle point we have to be concerned with is that the first touching of with might occur at . Here we manage to rule this case out by a rather subtle analysis. In view of Remark 4.1 below, We only need to consider the 3-dimensional case, i.e., , in which case is a line.
In the following we use to denote the outward unit normal of , to denote the tangent plane of at , and to denote a unit vector generating the line . Recall that is the outward unit normal of ,
Case 1. for some .
Without loss of generality, we assume is parallel to . Hence the sphere touches the plane only at the point . Since and are not parallel, then is not parallel to , thus the intersection of with must be a circle. Since , we know that touches only at . Hence is the tangential vector of at . Since is the first touching point of with from the interior, we see that is also the tangential vector of at (see Figure 4).

It follows that is tangent to at . We are now in the same situation as in Case 2, Theorem 1.1 or Theorem 1.4. Hence the contact angle of with at must be smaller than , which is a contradiction, since the contact angle of with is by assumption.
Case 2. Neither is parallel to , i.e., for .
In this case, must be circles.
Case 2.1. .
Since is not parallel to , there exists exactly one of , say , which does not coincide with . Then is not tangent to at and is a curve near . Since and , we see that is the tangential vector of . Since is the first touching point of with , is also the tangential vector of . Hence is tangent to at . We are again in the position as in Case 2, Theorem 1.1 or Theorem 1.4, and consequently we get a contradiction.
Case 2.2. .
In this case, , , cannot be parallel to . For simplicity of notation, in the following we use to indicate , respectively, and we adopt the notation
Let be the unit tangent vector of at such that for and be the unit tangent vector of at such that for . Since and are perpendicular to , we know that is parallel to . Similarly, since and are perpendicular to , we have that is parallel to .
Without loss of generality, we may assume that the origin and . Let be the unit tangent vector . With such choice of , we claim that and . Indeed, recall that by capillarity, is expressed as
Here is the outward unit normal vector of in at . Since lies on the same side of the line in , it yields that , thus we get
We will show that . Since touches from the interior at , we have for . Combining with , it follows that is pointing outward of at , see Figure 5, thus we obtain .

Note that and are perpendicular to . Without loss of generality, we may assume
A direct computation then yields
which implies the following fact: the two vectors and are in the opposite direction. Therefore,
By a similar argument, we obtain
Thanks to the fact that is the first touching point, we must have (see Figure 6)
(37) |

Let be such that for each . We can carry out the following computations.
Similarly,
Plugging into (37), we thus obtain
(38) |
(39) |
A crucial observation is that (38) is equivalent to (see the end of Appendix A)
(40) |
On the one hand, we have
(41) |
Since and , we have: and . It follows from (4) that
By our angle assumption , we get
which implies
(42) |
On the other hand, since , by Lemma A.2, we see that , where , satisfying . In view of Lemma A.1, it implies that
(43) |
From (42) and (43), we know that
Combining with (4), we find
(44) |
By a similar argument we obtain from (39) that
(45) |
and
(46) |
We may assume without loss of generality, then it follows from (45) that
Therefore, back to (44), we deduce that
(47) |
In the meanwhile, from (46), we deduce
(48) |
In conclusion, we have showed that the first touching point cannot occur at any . The rest of the proof follows similarly from Theorem 1.4. ∎
Remark 4.1.
We make a remark for the case of higher dimensions . In this case, is a -plane and is a closed -submanifold, playing the role as or . By letting be the unique unit vector in which is orthogonal to , and be the tangent vector of at which is orthogonal to , and be the tangent vector of at which is orthogonal to ( because is the first touching point), we may reduce the problem to the -dimensional case. In other words, all the vectors at we considered in the proof lie in the orthogonal -subspace of .
5. Alexandrov-Type Theorem
It is well-known that there exists at least one elliptic point for a closed embedded hypersurface in . Recall that an elliptic point is a point at which the principal curvatures are all positive with respect to the outward unit normal. We shall show that this fact is true for capillary hypersurfaces in the half-space, or in a wedge whenever condition (8) holds. Recall that for a wedge case, the existence of an elliptic point is no longer true if without any restriction. An example is easy to be found among ring type capillary surfaces a wedge. See a figure in [McCuan97].
We first need the following lemma (compare to Remark 3.1).
Lemma 5.1.
Let be a generalized wedge and . If , then for any , the sphere intersects at angle . In particular, in , the sphere intersects at angle .
Proof.
First, if , it is easy to see that for any , intersects . The outward unit normal to at is given by
Since , we see that
∎
Proposition 5.2.
Let be a smooth, compact, embedded -capillary hypersurface. Then there exists at least one elliptic point on .
Proof.
Let be the enclosed region of and , , we fix a point . Consider the family of the open balls . By Lemma 5.1, intersects with at the angle . Since is compact, for large enough, . Hence we can find the smallest , say , such that touches at a first time at some . For simplicity, we abbreviate by .
If , then and are tangent at . If , from the fact that both and intersect with at the angle , we conclude again that and are tangent at .
In both cases, we have that the principal curvatures of at are bigger than or equal to , which implies that is an elliptic point. ∎
Proposition 5.3.
Let be a classical wedge. Let be a smooth, compact, embedded -capillary surface. If , then there exists at least one elliptic point on .
Proof.
In view of Remark 4.1, we need only consider -dimensional case.
Since the corners () are smooth co-dimension two submanifolds in , is embedded in (see Section 2), when , we must have that is also a co-dimension two submanifold in . This means it has non-trivial interior relative to the topology of , therefore we are free to choose any in the interior of as the initial center of the sphere foliation . Thanks to Lemma A.4, any such foliation must have a first touching point with from outside as decreases from .
We consider the sphere which touches for the first time at some . Following the argument in LABEL:Prop-ellipticpoint1, we see that, if or , then must be an elliptic point of . Therefore, it remains to consider the case .
Since , we find
This implies that either (a) , or (b) , where is a unit vector perpendicular to both and , and is a non-zero constant.
We claim that the situation (b) does not occur. Otherwise, by the fact that and are unit vectors, we can deduce that . Thus we have
(49) |
Note that and , by the fact that and are outward normal vectors, we have
Since , is parallel to . It follows that and have the same sign, which contradicts (49) and concludes the claim.
From the claim, we see that only the situation (a) happens. This means, and are tangent at . Thus the principal curvatures of at are bigger than or equal to , which implies is an elliptic point. ∎
In view of the proof of LABEL:Prop-ellipticpoint2, the only place where we used the condition is that we have to use it in Lemma A.4, as , to conclude that the sphere foliation must touch . In this regard, we can remove the extra assumption by strengthening the angle condition to be . Indeed, we have the following
Proposition 5.4.
Let be a classical wedge. Let be a smooth, compact, embedded -capillary hypersurface with . Then there exists at least one elliptic point on .
If the hypersurfaces are CMC hypersurfaces, we have
Proposition 5.5.
Let be a classical wedge. Let be a smooth, compact, embedded -capillary hypersurface. If is of constant mean curvature , then , provided
Proof.
We remark that the free boundary case, that is has been proved by Lopez [Lopez14]. We consider here the general case.
To begin, we take to be the point of maximal distance of from the edge . Assume without loss of generality that . Let be the family of planes parallel to the edge of and having contact angle with . Starting from one of such a plane near infinite and moving it among this family until the first time that one of plane in touches . By definition of and , it is only possible that the first touching occurs at certain interior point of , at or at some .
For the former two cases, one can use the strong maximum principle for the interior point and the Hopf-lemma for the boundary point, and obtain . For the last case, we shall use our angle assumption .
Indeed, if , then the first touching point must not occur at any points of : since is a plane having contact angle with , we know that the contact angle of with , say , is . By the angle assumption, we find
(50) |
which is not possible if touches from outside for the first time.
If , then from the above discussion, and we can use the Hopf’s boundary point lemma again to find that . This completes the proof. ∎
Now, we are in the position to prove the Alexandrov-type theorem and the non-existence theorem.
Proof of Theorem 1.7.
Before we proceed the proof, we emphasize that the condition ensures the validity of the inequality
pointwisely on . In particular, as , we have
along .
On one hand, by virtue of LABEL:Prop-ellipticpoint2 and Gärding’s argument [Garding59] (see also [Ros87, Section 3]), we know that are positive, for . Applying Theorem 1.5 and using the Maclaurin inequality
we find
(51) |
and equality holds if and only if is a -capillary spherical cap.
Proof of Theorem 1.8.
For the CMC case, we notice that our condition (8) implies automatically, thanks to Lemma A.1. Therefore, we can use Proposition 5.5 to see that is of positive constant mean curvature.
For the constant case, since we assume , we may use Proposition 5.4 to conclude that has an elliptic point, from which we see that is a positive constant.
In view of this, arguing as in the proof of Theorem 1.7, we can use Proposition 2.3 together with Theorem 1.4 to show that the equality case happens in the Heintze-Karcher inequality (10), and hence must be a -spherical cap. However, if this is the case, a simple geometric relation (see Figure 7) then implies that , i.e., , which is incompatible with (11). The proof is complete. ∎

Appendix A Miscellaneous Results in Wedge
Lemma A.1.
Proof.
First we note that (11) can be viewed of somewhat a compatible condition for admitting spherical caps in a wedge(a wedge is determined by its dihedral angle .) with prescribed angles .
The following observation is important for our analysis in Theorem 1.5.
Lemma A.2.
Let be a classical wedge. If is a smooth, compact, embedded -capillary hypersurface. Let be given by such that If , then we have on . Moreover, for any , if and only if .
Proof.
In view of Remark 4.1, we need only consider the -dimensional case.
Let , In the following, we compute at . We have
Thus
Since , we deduce
which implies .
If , we have . Then we get
which is equivalent to . ∎
We point out that, in the discussion above, if intersects and transversally at , then Lemma A.2 is included in [Li21, Lemma 2.5].
Lemma A.3.
Let be a classical wedge. If is a smooth, compact, embedded -capillary with , then for any , the family of spheres must touch .
Proof.
The case follows trivially, since foliates the whole . As for , we proceed by the following observation.
Observation. Since for any , with . Moreover,
(52) |
In other words, the family of spheres foliates the half-space .
We claim that for any , there holds . To see this, we consider the following situations separately.
Case 1. .
By definition of , we see that with , away from the edge (see Figure 7 for illustration). Therefore for any , the open half-space determined by must intersect .
Case 2. .
In this case, since , we must have for any , according to Lemma A.2 (with chosen therein). However, this contradicts to our assumption on , precisely, that are smooth co-dimension two submanifolds in . This proves the claim and hence completes the proof.
∎
In the proof of LABEL:Prop-ellipticpoint2, we use the family of spheres for some in the interior of , the following observation is needed.
Lemma A.4.
Let be a classical wedge. If is a smooth, compact, embedded -capillary such that , then for any in the interior of , the family of spheres must touch .
Proof.
In view of Remark 4.1, we need only consider the -dimensional case.
By virtue of Lemma A.2, we have . The case follows trivially from Lemma A.3. As for , by Observation above and the fact that due to , it suffices to show that .
Claim. In (7), we have , provided that and
If the claim holds, a direct computation yields: for any ,
Notice that forms an orthonormal basis of , where is a fixed unit vector, parallel to , is the unit inwards pointing conormal of in . In this coordinate, since , it can be expressed as with . It follows that . Similarly, we have: for any , there holds , which implies that and proves the Lemma.
We are thus left to prove the Claim. Indeed, by Lemma A.2, let , then we have . In view of this, we obtain: , where is the unit outward normal of at . Due to the contact angle condition, we must have
(53) |
comparing with the definition of (7), we thus find:
Since is the outward unit normal of at , for any , we have
Indeed, it follows from and the fact is orthogonal to and is pointing outward at .
Meanwhile, since , we definitely have
Recall that and hence for each , we thus obtain
Combining all above and invoking that , we thus find: . This proves the Claim and completes the proof.
∎