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

Heintze-Karcher inequality and
capillary hypersurfaces in a wedge

Xiaohan Jia School of Mathematics
Southeast University
211189, Nanjing, P.R. China
[email protected]
Guofang Wang Mathematisches Institut
Universität Freiburg
Ernst-Zermelo-Str.1, 79104, Freiburg, Germany
[email protected]
Chao Xia School of Mathematical Sciences
Xiamen University
361005, Xiamen, P.R. China
[email protected]
 and  Xuwen Zhang School of Mathematical Sciences
Xiamen University
361005, Xiamen, P.R. China
[email protected]
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.

This work is supported by the NSFC (Grant No. 11871406, 12271449, 12126102)

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 Σ\Sigma in n+1\mathbb{R}^{n+1} with boundary Σ\partial\Sigma on a support hypersurface Sn+1S\subset\mathbb{R}^{n+1} is a critical point of the following functional

|Σ|cosθ|Ω¯S||\Sigma|-\cos\theta|\bar{\Omega}\cap S|

among all compact hypersurfaces with boundary Σ\partial\Sigma on SS under a volume constraint. Here Ω\Omega is the bounded domain enclosed by Σ\Sigma and SS, and |Σ||\Sigma| is the nn-dimensional area of Σ\Sigma, and θ(0,π)\theta\in(0,\pi). Equivalently, a capillary hypersurface is a constant mean curvature (CMC) hypersurface with boundary which intersects the support hypersurface SS at a constant angle θ\theta. 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 θ=π2\theta=\frac{\pi}{2}, 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 Σ\Sigma, which encloses a bounded domain Ω\Omega, in an (n+1)(n+1)-dimensional Riemannian manifold of nonnegative Ricci curvature, Heintze-Karcher’s theorem reads as follows,

|Ω|Σ0c(p)(1H(p)nt)ndtdA.\displaystyle|\Omega|\leq\int_{\Sigma}\int_{0}^{c(p)}\left(1-\frac{H(p)}{n}t\right)^{n}{\rm d}t{\rm d}A. (1)

Here H(p)H(p) is the mean curvature of Σ\Sigma at pp and c(p)c(p) is the length to reach the first focal point of Σ\Sigma from pp by the normal exponential map. As a direct consequence of (1), one deduces that

|Ω|nn+1Σ1HdA,\displaystyle|\Omega|\leq\frac{n}{n+1}\int_{\Sigma}\frac{1}{H}{\rm d}A, (2)

provided that Σ\Sigma is strictly mean convex, namely, H>0H>0 on Σ\Sigma. 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 n+1\mathbb{R}^{n+1} 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 n+1\mathbb{R}^{n+1} 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 +n+1:={xn+1:x,En+1>0}\mathbb{R}^{n+1}_{+}:=\{x\in\mathbb{R}^{n+1}:\langle x,E_{n+1}\rangle>0\}, where En+1=(0,,0,1)E_{n+1}=(0,\cdots,0,1), and for an embedded, compact, strictly mean-convex hypersurface Σ+n+1¯\Sigma\subset\overline{\mathbb{R}^{n+1}_{+}} with capillary boundary with a constant contact angle θ0(0,π2]\theta_{0}\in(0,\frac{\pi}{2}], there holds

Σ1HdAn+1n|Ω|+cosθ0(Σν,En+1dA)2ΣHν,En+1dA,\displaystyle\int_{\Sigma}\frac{1}{H}{\rm d}A\geq\frac{n+1}{n}|\Omega|+\cos\theta_{0}\frac{\left(\int_{\Sigma}\langle\nu,E_{n+1}\rangle{\rm d}A\right)^{2}}{\int_{\Sigma}H\langle\nu,E_{n+1}\rangle{\rm d}A}, (3)

with equality if and only if Σ\Sigma 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

Σn(1cosθ0ν,En+1)Hx,νdA=0,\displaystyle\int_{\Sigma}n(1-\cos\theta_{0}\langle\nu,E_{n+1}\rangle)-H\langle x,\nu\rangle{\rm d}A=0, (4)

a possible best form, which was conjectured in [JXZ22], is

Σ1cosθ0ν,En+1HdAn+1n|Ω|.\displaystyle\int_{\Sigma}\frac{1-\cos\theta_{0}\left<\nu,E_{n+1}\right>}{H}{\rm d}A\geq\frac{n+1}{n}|\Omega|. (5)

It is clear by using the Cauchy-Schwarz inequality that inequality (5) implies (3), provided that ν,En+1\langle\nu,E_{n+1}\rangle 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 θ0(0,π)\theta_{0}\in(0,\pi). Let Σ\Sigma be a hypersurface in +n+1¯\overline{\mathbb{R}^{n+1}_{+}} with (possibly non-connected) boundary Σ+n+1\partial\Sigma\subset\partial\mathbb{R}^{n+1}_{+}. The hypersurface intersects with the supported hyperplane +n+1\partial\mathbb{R}^{n+1}_{+} 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 θ0(0,π)\theta_{0}\in(0,\pi) and let Σ+n+1¯\Sigma\subset\overline{\mathbb{R}^{n+1}_{+}} be a smooth, compact, embedded, strictly mean convex θ\theta-capillary hypersurface, with θ(x)θ0\theta(x)\leq\theta_{0} for every xΣx\in\partial\Sigma. Let Ω\Omega denote the enclosed domain by Σ\Sigma and +n+1\partial\mathbb{R}^{n+1}_{+}. Then it holds

Σ1cosθ0ν,En+1HdAn+1n|Ω|.\displaystyle\int_{\Sigma}\frac{1-\cos\theta_{0}\left<\nu,E_{n+1}\right>}{H}{\rm d}A\geq\frac{n+1}{n}|\Omega|. (6)

Equality in (6) holds if and only if Σ\Sigma is a θ0\theta_{0}-capillary spherical cap.

Note that we also have removed the restriction that θ0π2\theta_{0}\leq\frac{\pi}{2}, comparing with the previous work [JXZ22]. As a direct application, we get the Alexandrov-type theorem for embedded capillary hypersurfaces in +n+1¯\overline{\mathbb{R}^{n+1}_{+}} with constant rr-th mean curvature, for any r{1,,n}r\in\{1,\cdots,n\}.

Corollary 1.2.

Let θ0(0,π)\theta_{0}\in(0,\pi) and r{1,,n}r\in\{1,\cdots,n\}. Let Σ+n+1¯\Sigma\subset\overline{\mathbb{R}^{n+1}_{+}} be a smooth, embedded, compact, θ0\theta_{0}-capillary hypersurface with constant rr-th mean curvature. Then Σ\Sigma is a θ0\theta_{0}-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 ζ(Σ,t)\zeta(\Sigma,t) defined in (33). To prove Theorem 1.1, we need to show the surjectivity of ζ\zeta onto the enclosed domain Ω\Omega, 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 Σ\Sigma be a closed, strictly mean convex hypersurface in n+1{\mathbb{R}}^{n+1} with enclosed domain Ω\Omega. Then it holds

Σ1HdAmaxe𝕊n|Σv,eHdA|n+1n|Ω|.\displaystyle\int_{\Sigma}\frac{1}{H}{\rm d}A-\max_{e\in{\mathbb{S}}^{n}}\left|\int_{\Sigma}\frac{\langle v,e\rangle}{H}{\rm d}A\right|\geq\frac{n+1}{n}|\Omega|.

Equality in (1.1) holds if and only if Σ\Sigma 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 α\alpha, which is also called an opening angle, lying in (0,π)(0,\pi). 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 𝐖\mathbf{W} be the unbounded closed region in n+1(n2)\mathbb{R}^{n+1}(n\geq 2), which are determined by finite many mutually intersecting hyperplanes P1,,PLP_{1},\ldots,P_{L}, for some integer 1Ln+11\leq L\leq n+1, such that the dihedral angle between PiP_{i} and PjP_{j}, iji\neq j, lies in (0,π)(0,\pi). We call such 𝐖\mathbf{W} a generalized wedge. PiPj,ijP_{i}\cap P_{j},i\neq j is called an edge of the wedge 𝐖\mathbf{W}. If L=2L=2, we call 𝐖\mathbf{W} a classical wedge. Let N¯i\bar{N}_{i} be the outwards pointing unit normal to PiP_{i} in 𝐖\mathbf{W} for i=1,,Li=1,\cdots,L. Thus {N¯1,,N¯L}\{\bar{N}_{1},\ldots,\bar{N}_{L}\} are linearly independent. Up to a translation, we may assume that the origin Oi=1LPiO\in\bigcap\limits_{i=1}^{L}P_{i}. Given θ0=(θ01,,θ0L)i=1L(0,π)\vec{\theta}_{0}=(\theta^{1}_{0},\cdots,\theta^{L}_{0})\in\prod\limits_{i=1}^{L}(0,\pi). Now we define an important vector 𝐤0\mathbf{k}_{0} associated with 𝐖\mathbf{W} and θ0\vec{\theta}_{0} by

𝐤0=i=1LciN¯i,\displaystyle\mathbf{k}_{0}=\sum_{i=1}^{L}c_{i}\bar{N}_{i}, (7)

where cic_{i} is such that 𝐤0,N¯i=cosθ0i\langle\mathbf{k}_{0},\bar{N}_{i}\rangle=\cos\theta_{0}^{i}. We say that Σ\Sigma is a θ\vec{\theta}-capillary hypersurface in 𝐖\mathbf{W} with θ=(θ1,θ2,,θL)\vec{\theta}=(\theta^{1},\theta^{2},\cdots,\theta^{L}) if it intersects 𝐖\partial\mathbf{W} at contact angle θi(x)\theta^{i}(x) for xΣPix\in\partial\Sigma\cap P_{i}. It is easy to see that 𝐤0\mathbf{k}_{0} is the center of the θ0\vec{\theta}_{0}-capillary spherical cap with radius 11. The following key assumption (8) has a clear geometric meaning that the unit sphere centered at 𝐤0\mathbf{k}_{0} intersects the edge of the wedge 𝐖\mathbf{W}.

We shall prove the following Heintze-Karcher-type inequality in a wedge.

Theorem 1.4.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a generalized wedge whose boundary consists of LL mutually intersecting closed hyperplanes {Pi}i=1L\{P_{i}\}_{i=1}^{L} and θ0i=1L(0,π)\vec{\theta}_{0}\in\prod\limits_{i=1}^{L}(0,\pi). Assume that

|𝐤0|1.\displaystyle|\mathbf{k}_{0}|\leq 1. (8)

Let Σ𝐖\Sigma\subset\mathbf{W} be a smooth, compact, embedded, strictly mean convex θ\vec{\theta}-capillary hypersurface with θi(x)θ0i\theta^{i}(x)\leq\theta_{0}^{i} for xΣPix\in\partial\Sigma\cap P_{i}, i=1,,Li=1,\ldots,L. Let Ω\Omega be the enclosed domain by Σ\Sigma and 𝐖\partial\mathbf{W}. Assume in addition that Σ\Sigma does not hit the edges of 𝐖\mathbf{W}, i.e.,

Σ(PiPj)=,ij.\displaystyle\Sigma\cap(P_{i}\cap P_{j})=\emptyset,\quad i\neq j. (9)

Then

Σ1+ν,𝐤0HdAn+1n|Ω|,\displaystyle\int_{\Sigma}\frac{1+\left<\nu,\mathbf{k}_{0}\right>}{H}{\rm d}A\geq\frac{n+1}{n}|\Omega|, (10)

with equality if and only if Σ\Sigma is a θ0\vec{\theta}_{0}-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 ζ(Σ,t)\zeta(\Sigma,t) which relates 𝐤0\mathbf{k}_{0}. To show the surjectivity of ζ\zeta, 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., L=2L=2. Precisely, we have the following

Theorem 1.5.

When L=2L=2, 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, L=2L=2 and θ0=(π2,π2)\vec{\theta}_{0}=(\frac{\pi}{2},\frac{\pi}{2}), i.e., Σ\Sigma is a free boundary hypersurface, for which 𝐤0=0,\mathbf{k}_{0}=0, (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 L>2L>2 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.
  • (i)

    In view of the Heintze-Karcher inequality (10) we have established, (8) could not be removed.

  • (ii)

    When L=1L=1, it is nothing but the half-space case and (8) is satisfied automatically. When L=2L=2, (8) is equivalent to

    |π(θ01+θ02)|απ|θ01θ02|.\displaystyle|\pi-(\theta_{0}^{1}+\theta_{0}^{2})|\leq\alpha\leq\pi-|\theta_{0}^{1}-\theta_{0}^{2}|. (11)

    where α(0,π)\alpha\in(0,\pi) is the opening angle of the wedge, or the dihedral angle between P1P_{1} and P2P_{2}, see Lemma A.1. Similarly, |𝐤0|<1|\mathbf{k}_{0}|<1 is equivalent to (11) with strict inequalities.

  • (iii)

    By virtue of Lemma A.2, Condition (8) is satisfied, provided there exists a θ0\vec{\theta}_{0}-capillary hypersurface Σ\Sigma in 𝐖\mathbf{W} which satisfies ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset.

  • (iv)

    Condition (8) is also closely related to the existence of elliptic points of capillary hypersurfaces. See Section 5 below.

  • (v)

    Condition (11) appears also in the regularity of capillary surfaces at corner. See the last paragraph of the Introduction.

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 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge whose boundary consists of P1,P2P_{1},P_{2} with θ0i=12(0,π)\vec{\theta}_{0}\in\prod\limits_{i=1}^{2}(0,\pi). Let Σ𝐖\Sigma\subset\mathbf{W} be a smooth, compact and embedded θ0\vec{\theta}_{0}-capillary hypersurface with constant rr-mean curvature, r{1,,n}r\in\{1,\cdots,n\}. Assume ΣP1P2\Sigma\cap P_{1}\cap P_{2}\not=\emptyset. Then Σ\Sigma is a θ0\vec{\theta}_{0}-capillary spherical cap.

Theorem 1.8.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge whose boundary consists of P1,P2P_{1},P_{2} with θ0i=12(0,π)\vec{\theta}_{0}\in\prod\limits_{i=1}^{2}(0,\pi). Then there exists no smooth, compact and embedded θ0\vec{\theta}_{0}-capillary, CMC hypersurface such that ΣP1P2=\Sigma\cap P_{1}\cap P_{2}=\emptyset and |𝐤0|1|\mathbf{k}_{0}|\leq 1. Moreover, there exists no smooth, compact, embedded, θ0\vec{\theta}_{0}-capillary hypersurface of constant rr-mean curvature for some r{2,,n}r\in\{2,\cdots,n\}, such that ΣP1P2=\Sigma\cap P_{1}\cap P_{2}=\emptyset and |𝐤0|<1|\mathbf{k}_{0}|<1.

As a consequence of Theorem 1.7, Theorem 1.8 and Remark 1.6 (iii), we have the following

Theorem 1.9.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge whose boundary consists of P1,P2P_{1},P_{2} with θ0i=12(0,π)\vec{\theta}_{0}\in\prod\limits_{i=1}^{2}(0,\pi). Let Σ𝐖\Sigma\subset\mathbf{W} be a smooth, compact and embedded θ0\vec{\theta}_{0}-capillary CMC hypersurface. Then Σ\Sigma is a θ0\vec{\theta}_{0}-capillary spherical cap which intersects with the edge P1P2P_{1}\cap P_{2} if and only if |𝐤0|1|\mathbf{k}_{0}|\leq 1.

Several remarks and questions are in order.

Remark 1.10.
  • (i)

    McCuan [McCuan97, Theorem 2] proved that any θ0\vec{\theta}_{0}-capillary spherical cap which is disjoint with the edge must satisfy θ01+θ02>π+α\theta_{0}^{1}+\theta_{0}^{2}>\pi+\alpha. This condition implies that |𝐤0|>1|\mathbf{k}_{0}|>1, see Lemma A.1.

  • (ii)

    Lopez [Lopez14] proved an Alexandrov-type theorem for embedded CMC capillary surfaces with θ0=(π2,π2)\vec{\theta}_{0}=(\frac{\pi}{2},\frac{\pi}{2}), i.e., the free boundary case. Note that in this case, |𝐤0|1|\mathbf{k}_{0}|\leq 1 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 θ0\vec{\theta}_{0}-hypersurfaces for a general θ0\vec{\theta}_{0}.

  • (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

    θ01+θ02π+α,\displaystyle\theta_{0}^{1}+\theta_{0}^{2}\leq\pi+\alpha, (12)

    by developing spherical reflection technique, when n=2n=2. Note that the angle relation (12) is weaker than |𝐤0|1|\mathbf{k}_{0}|\leq 1, 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 Σ\partial\Sigma for n=2n=2 or the convexity of Σ\partial\Sigma for n3n\geq 3. 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 ff over a domain, say Ω\Omega, 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 Ω\Omega 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 ff 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 α>π|θ01θ02|\alpha>\pi-|\theta^{1}_{0}-\theta^{2}_{0}|, they conjectured that ff 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 Ω\Omega be the bounded domain in 𝐖\mathbf{W} with piecewise smooth boundary Ω=Σ(i=1LTi)\partial\Omega=\Sigma\cup(\bigcup_{i=1}^{L}T_{i}), where Σ=Ω𝐖̊¯\Sigma=\overline{\partial\Omega\cap\mathring{\mathbf{W}}} is a smooth compact embedded θ\vec{\theta}-capillary hypersurface in 𝐖\mathbf{W} and Ti=ΩPiT_{i}=\partial\Omega\cap P_{i}. Denote the corners by Γi=ΣTi\Gamma_{i}=\Sigma\cap T_{i}, which are smooth, co-dimension two submanifolds in n+1\mathbb{R}^{n+1}. For the sake of simplicity, we denote by Γ\Gamma the union of Γi\Gamma_{i}, i.e., Γ=i=1LΓi\Gamma=\bigcup_{i=1}^{L}\Gamma_{i}. We use the following notation for normal vector fields. Let ν\nu and N¯i\bar{N}_{i} be the outward unit normal to Σ\Sigma and PiP_{i} (with respect to Ω\Omega) respectively. Let μi\mu_{i} be the outward unit co-normal to Γi=ΣPiΣ\Gamma_{i}=\partial\Sigma\cap P_{i}\subset\Sigma and ν¯i\bar{\nu}_{i} be the outward unit co-normal to ΓiTi\Gamma_{i}\subset T_{i}. Under this convention, along each Γi\Gamma_{i} {ν,μi}\{\nu,\mu_{i}\} and {ν¯i,N¯i}\{\bar{\nu}_{i},\bar{N}_{i}\} span the same 2-dimensional plane and have the same orientation in the normal bundle of Σn+1\partial\Sigma\subset\mathbb{R}^{n+1}. Hence one can define the contact angle function along each Γi\Gamma_{i}, θi:Γi(0,π)\theta_{i}:\Gamma_{i}\to(0,\pi) by

μi(x)=sinθi(x)N¯i+cosθi(x)ν¯i(x),\displaystyle\mu_{i}(x)=\sin{\theta^{i}(x)}\bar{N}_{i}+\cos\theta^{i}(x)\bar{\nu}_{i}(x), (13)
ν(x)=cosθi(x)N¯i+sinθi(x)ν¯i(x).\displaystyle\nu(x)=-\cos\theta^{i}(x)\bar{N}_{i}+\sin\theta^{i}(x)\bar{\nu}_{i}(x). (14)

Let θ(x)\vec{\theta}(x) denote the LL-tuple (θ1(x),,θL(x))(\theta^{1}(x),\ldots,\theta^{L}(x)). We call Σ\Sigma a θ\vec{\theta}-capillary hypersurface, if we want to emphasize the contact angle function. We also use θ0\vec{\theta}_{0}-capillary hypersurface to denote such a hypersurface with θθ0\vec{\theta}\equiv\vec{\theta}_{0}, a vector θ0i=1L(0,π).\vec{\theta}_{0}\in\prod_{i=1}^{L}(0,\pi).

We denote by ¯\bar{\nabla}, Δ¯\bar{\Delta}, ¯2\bar{\nabla}^{2} and div¯\bar{\rm div}, the gradient, the Laplacian, the Hessian and the divergence on n+1\mathbb{R}^{n+1} respectively, while by \nabla, Δ\Delta, 2\nabla^{2} and div{\rm div}, the gradient, the Laplacian, the Hessian and the divergence on the smooth part of Ω\partial\Omega, respectively. Let gg, hh and HH be the first, second fundamental forms and the mean curvature of the smooth part of Ω\partial\Omega respectively. Precisely, h(X,Y)=¯Xν,Yh(X,Y)=\langle\bar{\nabla}_{X}\nu,Y\rangle and H=trg(h)H={\rm tr}_{g}(h). In particular, since PiP_{i} is planar, the second fundamental form hi0h_{i}\equiv 0, correspondingly, the mean curvature HiH_{i} vanishes.

We need the following structural lemma for compact hypersurfaces in n+1\mathbb{R}^{n+1} with boundary, which is well-known and widely used, see [AS16, JXZ22].

Lemma 2.1.

Let Σn+1\Sigma\subset\mathbb{R}^{n+1} be a smooth compact hypersurface with boundary. Then it holds that

nΣνdA=Σ{x,μνx,νμ}ds.\displaystyle n\int_{\Sigma}\nu{\rm d}A=\int_{\partial\Sigma}\left\{\left<x,\mu\right>\nu-\left<x,\nu\right>\mu\right\}{\rm d}s. (15)
Proof.

Let Z=ν,exTx,νeTZ=\left<\nu,e\right>x^{T}-\left<x,\nu\right>e^{T} for any constant vector ee, where xTx^{T} and eTe^{T} denote the tangential component of xx and ee respectively. One computes that

div(Z)=nν,e.{\rm div}(Z)=n\left<\nu,e\right>.

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 n+1\mathbb{R}^{n+1}, see e.g., [AS16, LX17, WX19].

Lemma 2.2.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a wedge and θ0i(0,π)\theta^{i}_{0}\in(0,\pi) for i=1,,Li=1,\ldots,L. If Σ𝐖\Sigma\subset\mathbf{W}is a smooth θ0\vec{\theta}_{0}-capillary hypersurface, then along Σ\partial\Sigma, μi\mu_{i} is a principal direction of Σ\Sigma.

Proof.

For the completeness we provide a proof. It suffices to prove that h(μi,X)=0h(\mu_{i},X)=0 for any vector XX tangent to Σ\partial\Sigma. Indeed,

h(μi,X)=\displaystyle h(\mu_{i},X)= ¯Xμi,ν=¯X(sinθiN¯i+cosθiν¯i),cosθiN¯i+sinθiν¯i\displaystyle\left<\bar{\nabla}_{X}\mu_{i},\nu\right>=\left<\bar{\nabla}_{X}\left(\sin\theta^{i}\bar{N}_{i}+\cos\theta^{i}\bar{\nu}_{i}\right),-\cos\theta^{i}\bar{N}_{i}+\sin\theta^{i}\bar{\nu}_{i}\right>
=\displaystyle= ¯Xν¯i,N¯i=hi(ν¯i,X)=0,\displaystyle\left<\bar{\nabla}_{X}\bar{\nu}_{i},\bar{N}_{i}\right>=-h_{i}(\bar{\nu}_{i},X)=0, (16)

where we have used (13), the constancy of θi\theta_{i}, the fact that ν¯i,N¯i\bar{\nu}_{i},\bar{N}_{i} are unit vector fields, and hi=0h_{i}=0 since PiP_{i} are totally geodesic. This completes the proof. ∎

The rr-th mean curvature HrH_{r} of Σ\Sigma is defined by the identity:

𝒫n(t)=i=1n(1+tκi)=i=0n(ni)Hiti\displaystyle\mathcal{P}_{n}(t)=\prod_{i=1}^{n}(1+t\kappa_{i})=\sum_{i=0}^{n}\binom{n}{i}H_{i}t^{i} (17)

for all real number tt. Thus H1=HnH_{1}=\frac{H}{n} is the mean curvature of Σ\Sigma and HnH_{n} is the Gaussian curvature, and we adopt the convention that H0=1H_{0}=1.

We have the following Minkowski-type formula for θ0\vec{\theta}_{0}-capillary hypersurfaces.

Proposition 2.3.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a wedge and Σ𝐖\Sigma\subset\mathbf{W} be a θ0\vec{\theta}_{0}-capillary hypersurface. Then it holds that for r=1,,nr=1,\ldots,n,

Σ(Hr1(1+ν,𝐤0)Hrx,ν)dA=0.\displaystyle\int_{\Sigma}\left(H_{r-1}\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)-H_{r}\left<x,\nu\right>\right){\rm d}A=0. (18)

In particular, if L=1L=1, i.e. Σ¯+n+1\Sigma\subset\bar{\mathbb{R}}^{n+1}_{+}, then

ΣHr1(1cosθ0ν,En+1)Hrx,νdA=0.\displaystyle\int_{\Sigma}H_{r-1}(1-\cos\theta_{0}\langle\nu,E_{n+1}\rangle)-H_{r}\langle x,\nu\rangle{\rm d}A=0. (19)
Proof.

The case for r=1r=1 has been proved in [LX17, Lemma 5]. For the sake of completeness, we include the proof here.

Since

div(xT)=nHx,ν,{\rm div}(x^{T})=n-H\left<x,\nu\right>,

by using integration by parts in Σ\Sigma, we get

Σ(nHx,ν)dA=i=1LΓix,μidsi.\displaystyle\int_{\Sigma}(n-H\langle x,\nu\rangle){\rm d}A=\sum_{i=1}^{L}\int_{\Gamma_{i}}\langle x,\mu_{i}\rangle{\rm d}s_{i}. (20)

From the capillary boundary condition (13) it is easy to see that on each Γi\Gamma_{i}

cosθ0ix,ν+sinθ0ix,μi=x,N¯i=0.\displaystyle-\cos\theta^{i}_{0}\left<x,\nu\right>+\sin\theta^{i}_{0}\left<x,\mu_{i}\right>=\left<x,\bar{N}_{i}\right>=0. (21)

By (15), (13) and (21), we get

nΣν,𝐤0dA\displaystyle n\int_{\Sigma}\langle\nu,\mathbf{k}_{0}\rangle{\rm d}A =i=1LΓi(x,μiν,𝐤0x,νμi,𝐤0)dsi\displaystyle=\sum_{i=1}^{L}\int_{\Gamma_{i}}\left(\left<x,\mu_{i}\right>\langle\nu,\mathbf{k}_{0}\rangle-\left<x,\nu\right>\langle\mu_{i},\mathbf{k}_{0}\rangle\right){\rm d}s_{i}
=i=1LΓi(x,μiν,𝐤0+x,νcosθ0isinθ0i(1+ν,𝐤0))dsi\displaystyle=\sum_{i=1}^{L}\int_{\Gamma_{i}}\left(\left<x,\mu_{i}\right>\left<\nu,\mathbf{k}_{0}\right>+\left<x,\nu\right>\frac{-\cos\theta^{i}_{0}}{\sin\theta^{i}_{0}}\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)\right){\rm d}s_{i}
=i=1LΓi(x,μiν,𝐤0x,μi(1+ν,𝐤0))dsi\displaystyle=\sum_{i=1}^{L}\int_{\Gamma_{i}}\left(\left<x,\mu_{i}\right>\left<\nu,\mathbf{k}_{0}\right>-\left<x,\mu_{i}\right>\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)\right){\rm d}s_{i}
=i=1LΓix,μidsi,\displaystyle=-\sum_{i=1}^{L}\int_{\Gamma_{i}}\left<x,\mu_{i}\right>{\rm d}s_{i}, (22)

It follows from (20) and (2) that

Σn(1+ν,𝐤0)Hx,νdA=0.\displaystyle\int_{\Sigma}n(1+\langle\nu,\mathbf{k}_{0}\rangle)-H\langle x,\nu\rangle{\rm d}A=0. (23)

Now we prove (18) for general rr. For a small real number t>0t>0, consider a family of hypersurfaces with boundary Σt\Sigma_{t}, defined by

y:=φt(x)=x+t(ν(x)+𝐤0)xΣ.y:=\varphi_{t}(x)=x+t(\nu(x)+\mathbf{k}_{0})\quad x\in\Sigma.

We claim that Σt\Sigma_{t} is also a θ0\vec{\theta}_{0}-capillary hypersurface in 𝐖\mathbf{W}. In fact, if e1,,ene_{1},\ldots,e_{n} are principal directions of a point of Σ\Sigma and κi\kappa_{i} are the corresponding principal curvatures, we have

(φt)(ei)=¯eiφt=(1+tκi)ei,i=1,,n.\displaystyle(\varphi_{t})_{\ast}(e_{i})=\bar{\nabla}_{e_{i}}\varphi_{t}=(1+t\kappa_{i})e_{i},\quad i=1,\ldots,n. (24)

From (24), we see that νt(y)=ν(x)\nu_{t}(y)=\nu(x), where νt(y)\nu_{t}(y) denotes the outward unit normal of Σt\Sigma_{t} at y=φt(x)y=\varphi_{t}(x). Moreover, the capillarity condition (14) implies: for any xΣPix\in\partial\Sigma\cap P_{i}, we have

ν(x)+𝐤0,N¯i=cosθ0i+cosθ0i=0,\displaystyle\left<\nu(x)+\mathbf{k}_{0},\bar{N}_{i}\right>=-\cos\theta_{0}^{i}+\cos\theta_{0}^{i}=0, (25)

in other words, φt(x)Pi\varphi_{t}(x)\in P_{i}, and hence Σtφt(Σ)\partial\Sigma_{t}\subset\varphi_{t}(\partial\Sigma). In view of this, we have: νt(y),N¯i=ν(x),N¯i=cosθ0i\left<\nu_{t}(y),\bar{N}_{i}\right>=\left<\nu(x),\bar{N}_{i}\right>=-\cos\theta^{i}_{0}; that is, Σt\Sigma_{t} is also a θ0\vec{\theta}_{0}-capillary hypersurface in 𝐖\mathbf{W}.

Therefore, we can exploit (23) to find that

Σt=φt(Σ)n(1+νt,𝐤0)H(t)yy,νtdAt(y)=0.\displaystyle\int_{\Sigma_{t}=\varphi_{t}(\Sigma)}n\left(1+\left<\nu_{t},\mathbf{k}_{0}\right>\right)-H(t)\mid_{y}\left<y,\nu_{t}\right>{\rm d}A_{t}(y)=0. (26)

By (24), the tangential Jacobian of φt\varphi_{t} along Σ\Sigma at xx is just

JΣφt(x)=i=1n(1+tκi(x))=𝒫n(t),\displaystyle{\rm J}^{\Sigma}\varphi_{t}(x)=\prod_{i=1}^{n}(1+t\kappa_{i}(x))=\mathcal{P}_{n}(t), (27)

where 𝒫n(t)\mathcal{P}_{n}(t) is the polynomial defined in (17). Moreover, using (24) again, we see that the corresponding principal curvatures are given by

κi(φt(x))=κi(x)1+tκi(x).\displaystyle\kappa_{i}(\varphi_{t}(x))=\frac{\kappa_{i}(x)}{1+t\kappa_{i}(x)}. (28)

Hence fix xΣx\in\Sigma, the mean curvature of Σt\Sigma_{t} at φt(x)\varphi_{t}(x), say H(t)H(t), is given by

H(t)=𝒫n(t)𝒫n(t)=i=0ni(ni)Hiti1𝒫n(t),\displaystyle H(t)=\frac{\mathcal{P}^{\prime}_{n}(t)}{\mathcal{P}_{n}(t)}=\frac{\sum_{i=0}^{n}i\binom{n}{i}H_{i}t^{i-1}}{\mathcal{P}_{n}(t)}, (29)

where Hi=Hi(x)H_{i}=H_{i}(x) is the ii-th mean curvature of Σ\Sigma at xx.

Using the area formula, (27) and (29), we find from (26) that

Σn(1+ν,𝐤0)𝒫n(t)t(1+ν,𝐤0)𝒫n(t)𝒫n(t)x,νdAx=0.\displaystyle\int_{\Sigma}n\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)\mathcal{P}_{n}(t)-t\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)\mathcal{P}^{\prime}_{n}(t)-\mathcal{P}^{\prime}_{n}(t)\left<x,\nu\right>{\rm d}A_{x}=0. (30)

As the left hand side in this equality is a polynomial in the time variable tt, this shows that all its coefficients vanish, and hence

Σ(1+ν,𝐤0)Hr1Hrx,νdA,r=1,,n.\displaystyle\int_{\Sigma}\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)H_{r-1}-H_{r}\left<x,\nu\right>{\rm d}A,\quad r=1,\ldots,n. (31)

This gives (18). ∎

We have also

(n+1)|Ω|=Σx,νdA,\displaystyle(n+1)|\Omega|=\int_{\Sigma}\left<x,\nu\right>{\rm d}A, (32)

which is easy to prove. If one views (32) as one of (18) with r=0r=0, the Heintze-Karcher inequality (6) that we want to prove could also be viewed one of them with r=1r=-1. 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, Σ\Sigma 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 Σ+n+1¯\Sigma\subset\overline{\mathbb{R}^{n+1}_{+}} be a θ\theta-capillary hypersurface with θθ0\theta\leq\theta_{0} along Σ\partial\Sigma. For any xΣx\in\Sigma, let {ei=ei(x)}\{e_{i}=e_{i}(x)\} be the set of unit principal vectors of Σ\Sigma at xx and {κi(x)}\{\kappa_{i}(x)\} the set of corresponding principal curvatures. Since Σ\Sigma is strictly mean convex,

maxiκi(x)H(x)n>0, for xΣ.\max_{i}{\kappa_{i}(x)}\geq\frac{H(x)}{n}>0,\hbox{ for }x\in\Sigma.

We define

Z={(x,t)Σ×:0<t1maxκi(x)},\displaystyle Z=\left\{(x,t)\in\Sigma\times\mathbb{R}:0<t\leq\frac{1}{\max{\kappa_{i}(x)}}\right\},

and

ζ:Zn+1,\displaystyle\zeta:Z\to\mathbb{R}^{n+1},
ζ(x,t)=xt(ν(x)cosθ0En+1).\displaystyle\zeta(x,t)=x-t\left(\nu(x)-\cos\theta_{0}E_{n+1}\right). (33)

ζ\zeta gives a family of hypersurfaces ζ(Σ,t)\zeta(\Sigma,t), which are the modified parallel hypersurfaces mentioned above.

Claim: Ωζ(Z)\Omega\subset\zeta(Z).

Indeed, let us denote by Br(x)B_{r}(x) the closed ball centered at xx of radius rr, and Sr(x)=Br(x)S_{r}(x)=\partial B_{r}(x). For any yΩy\in\Omega, we consider a family of spheres {Sr(yrcosθ0En+1)}r0\{S_{r}(y-r\cos\theta_{0}E_{n+1})\}_{r\geq 0}. Since yΩy\in\Omega is an interior point, when rr is small enough, we have Br(yrcosθ0En+1)ΩB_{r}(y-r\cos\theta_{0}E_{n+1})\subset\subset\Omega. Since |cosθ0|<1|\cos\theta_{0}|<1, it is easy to see that the spheres gives a foliation of n+1\mathbb{R}^{n+1}. Hence Sr(yrcosθ0En+1)S_{r}(y-r\cos\theta_{0}E_{n+1}) must touch Σ\Sigma as we increase the radius rr. As a conclusion, for any yΩy\in\Omega, there exists xΣx\in\Sigma and ry>0r_{y}>0, such that Sr(yrycosθ0En+1)S_{r}(y-r_{y}\cos\theta_{0}E_{n+1}) touches Σ\Sigma for the first time, at the point xΣx\in\Sigma. We have two cases.

Case 1. xΣ̊x\in\mathring{\Sigma}.

In this case, since xΣ̊x\in\mathring{\Sigma}, the sphere Sr(yrycosθ0En+1)S_{r}(y-r_{y}\cos\theta_{0}E_{n+1}) is tangent to Σ\Sigma at xx from the interior. It follows that ry1maxκi(x)r_{y}\leq\frac{1}{\max{\kappa_{i}(x)}}. Invoking the definition of ZZ and ζ\zeta, we find that yζ(Z)y\in\zeta(Z) in this case.

Case 2. xΣx\in\partial\Sigma.

We will rule out this case by the condition on the contact angle function θ\theta of Σ\Sigma. In this case, by the first touching property of xx, the contact angle θy\theta_{y} of Sr(yrycosθ0En+1)S_{r}(y-r_{y}\cos\theta_{0}E_{n+1}) with +n+1\partial\mathbb{R}^{n+1}_{+} is smaller than or equals to θ(x)\theta(x), which is smaller than or equals to θ0\theta_{0}, by assumption. (see Figure 1 for an illustration). However θyθ0\theta_{y}\leq\theta_{0} implies that y,En+1<0\langle y,E_{n+1}\rangle<0, a contradiction to yΩ+n+1y\in\Omega\subset\mathbb{R}^{n+1}_{+}. The Claim is thus proved.

Refer to caption
Figure 1. Boundary touching.

By a simple computation, we find

tζ(x,t)\displaystyle\partial_{t}\zeta(x,t) =(ν(x)cosθ0En+1),\displaystyle=-\left(\nu(x)-\cos\theta_{0}E_{n+1}\right),
¯eiζ(x,t)\displaystyle\bar{\nabla}_{e_{i}}\zeta(x,t) =(1tκi(x))ei.\displaystyle=\left(1-t\kappa_{i}(x)\right)e_{i}.

Hence the tangential Jacobian of ζ\zeta along ZZ, at (x,t)(x,t) is just

JZζ(x,t)=(1cosθ0ν,En+1)i=1n(1tκi).{\rm J}^{Z}\zeta(x,t)=(1-\cos\theta_{0}\langle\nu,E_{n+1}\rangle)\prod_{i=1}^{n}(1-t\kappa_{i}).

By virtue of the fact that Ωζ(Z)\Omega\subset\zeta(Z), the area formula yields

|Ω||ζ(Z)|\displaystyle|\Omega|\leq|\zeta(Z)|\leq ζ(Z)0(ζ1(y))dy=ZJZζdn+1\displaystyle\int_{\zeta(Z)}\mathcal{H}^{0}(\zeta^{-1}(y)){\rm d}y=\int_{Z}{\rm J}^{Z}\zeta{\rm d}\mathcal{H}^{n+1}
=\displaystyle= ΣdA01max{κi(x)}(1cosθ0ν,En+1)i=1n(1tκi(x))dt.\displaystyle\int_{\Sigma}{\rm d}A\int_{0}^{\frac{1}{\max\left\{\kappa_{i}(x)\right\}}}\left(1-\cos\theta_{0}\left<\nu,E_{n+1}\right>\right)\prod_{i=1}^{n}(1-t\kappa_{i}(x)){\rm d}t.

By the AM-GM inequality, 1cosθ0ν,En+1>01-\cos\theta_{0}\left<\nu,E_{n+1}\right>>0 on Σ\Sigma, and the fact that max{κi(x)}H(x)/n\max\left\{\kappa_{i}(x)\right\}\geq H(x)/n, we obtain

|Ω|\displaystyle|\Omega|\leq ΣdA01max{κi(x)}(1cosθ0ν,En+1)(1ni=1n(1tκi(x)))ndt\displaystyle\int_{\Sigma}{\rm d}A\int_{0}^{\frac{1}{\max\left\{\kappa_{i}(x)\right\}}}\left(1-\cos\theta_{0}\langle\nu,E_{n+1}\rangle\right)\left(\frac{1}{n}\sum_{i=1}^{n}\left(1-t\kappa_{i}(x)\right)\right)^{n}{\rm d}t
\displaystyle\leq Σ(1cosθ0ν,En+1)dA0nH(x)(1tH(x)n)ndt\displaystyle\int_{\Sigma}\left(1-\cos\theta_{0}\langle\nu,E_{n+1}\rangle\right){\rm d}A\int_{0}^{\frac{n}{H(x)}}\left(1-t\frac{H(x)}{n}\right)^{n}{\rm d}t
=\displaystyle= nn+1Σ(1cosθ0ν,En+1)HdA,\displaystyle\frac{n}{n+1}\int_{\Sigma}\frac{(1-\cos\theta_{0}\langle\nu,E_{n+1}\rangle)}{H}{\rm d}A,

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 Σ\Sigma, and it follows that Σ\Sigma is a spherical cap. Apparently, the contact angle of a spherical cap with a hyperplane is a constant, say θ\theta. It is easy to see that θ=θ0\theta=\theta_{0}, and hence Σ\Sigma must be a θ0\theta_{0}-capillary spherical cap. Conversely, when Σ\Sigma is a θ0\theta_{0}-capillary spherical cap, then HH is a positive constant. By virtue of the Minkowski formula (19) for r=1r=1, we see that equality in (6) holds. ∎

Let us close this section with a remark. In the proof of Ωζ(Z)\Omega\subset\zeta(Z), our choice of the touching balls is enlightened by the following observation.

Remark 3.1 (Foliation of θ0\theta_{0}-capillary hypersurfaces).

In [MR91], to prove the Heintze-Karcher inequality for closed hypersurfaces, one shall ‘sweepout’ the domain Ω\Omega by a foliation around any point pΩp\in\Omega, whose leaves are level-sets of the distance function to pp. The key point is that, such ‘sweep-outs’ coincides with the domain Ω\Omega, if and only if Ω\Omega is a ball and pp 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 θ0\theta_{0}-ball with the foliation, whose leaves are θ0\theta_{0}-spherical caps(as illustrated in Figure 2).

Refer to caption

(a) Sweepout of θ0\theta_{0}-domain.

Refer to caption

(b) Sweepout of generic domain.

Figure 2. Capillary foliation.

4. Heintze-Karcher Inequality in a Wedge

In this section, we prove Theorem 1.4 and Theorem 1.5 in a wedge 𝐖\mathbf{W}, following largely from the proof of the Heintze-Karcher-type inequality presented in the previous section.

Proof of Theorem 1.4.

Let Σ𝐖\Sigma\subset\mathbf{W} be a compact embedded hypersurface with θ(x)\vec{\theta}(x)-capillary boundary, where θi(x)θ0i\theta^{i}(x)\leq\theta^{i}_{0} for each ii and every xΓx\in\Gamma. As above, for any xΣx\in\Sigma, let {ei}\{e_{i}\} be the set of principal unit vectors of Σ\Sigma at xx and {κi(x)}\{\kappa_{i}(x)\} the set of the corresponding principal curvatures. Now we define modified parallel hypersurfaces by

Z={(x,t)Σ×:0<t1maxκi(x)},\displaystyle Z=\left\{(x,t)\in\Sigma\times\mathbb{R}:0<t\leq\frac{1}{\max{\kappa_{i}(x)}}\right\},
ζ(x,t)=xt(ν(x)+𝐤0), (x,t)Z.\displaystyle\zeta(x,t)=x-t\left(\nu(x)+\mathbf{k}_{0}\right),\text{ }(x,t)\in Z.

As in Theorem 1.1, we shall show that Ωζ(Z)\Omega\subset\zeta(Z). Let yΩy\in\Omega. We consider the sphere foliation {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0}. By virtue of Lemma A.3, there exists some ry>0r_{y}>0 such that Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) touches Σ\Sigma from the interior at a first touching point xΣx\in\Sigma.

Case 1. xΣ̊.x\in\mathring{\Sigma}. We can get yζ(Z)y\in\zeta(Z), as argued in Case 1, Theorem 1.1.

Case 2. xΣx\in\partial\Sigma. By assumption (9), xΣP̊ix\in\partial\Sigma\cap\mathring{P}_{i} for some ii. We will rule out this case again by virtue of the capillarity of Σ\Sigma. In this case, Sry(y+ry𝐤0)PiS_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{i} is tangent to ΣPi\Sigma\cap P_{i}, and the touching angle of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with PiP_{i} at xΓix\in\Gamma_{i} must be smaller than θi(x)\theta^{i}(x), and it follows from the geometric relation that yy lies outside 𝐖\mathbf{W}. Precisely, up to a rotation, we may assume that the touching plane is {xn+1=0}\left\{x_{n+1}=0\right\}, say P1P_{1}, and we denote by θ¯1(x)\bar{\theta}^{1}(x) the touching angle, satisfying θ¯1(x)θ1(x)\bar{\theta}^{1}(x)\leq\theta^{1}(x), due to the first touch. From the geometric relation(see Figure 3), we find, y+ry(𝐤0cosθ¯1(x)N¯1)P1y+r_{y}\left(\mathbf{k}_{0}-\cos\bar{\theta}^{1}(x)\bar{N}_{1}\right)\in P_{1}, the angle relation θ¯1(x)θ1(x)θ01<π\bar{\theta}^{1}(x)\leq\theta^{1}(x)\leq\theta^{1}_{0}<\pi then implies that yn+1𝐖y\in\mathbb{R}^{n+1}\setminus\mathbf{W} 222Notice that 𝐤0,N¯1=cosθ01\left<\mathbf{k}_{0},\bar{N}_{1}\right>=\cos\theta_{0}^{1}, which means moving along 𝐤0\mathbf{k}_{0} with distance ryr_{y} is indeed moving along N¯1\bar{N}_{1} with distance rycosθ1r_{y}\cos\theta^{1}., which contradicts to the fact that yΩ𝐖y\in\Omega\subset\mathbf{W}. Therefore, we complete the proof that Ωζ(Z)\Omega\subset\zeta(Z).

Refer to caption
Figure 3. Touching supporting hyperplanes in the interior.

By a simple computation as Theorem 1.1, we see, the tangential Jacobian of ζ\zeta along ZZ at (x,t)(x,t) is just

JZζ(x,t)=(1+ν,𝐤0)i=1n(1tκi).{\rm J}^{Z}\zeta(x,t)=(1+\langle\nu,\mathbf{k}_{0}\rangle)\prod_{i=1}^{n}(1-t\kappa_{i}).

By a similar argument as Theorem 1.1, we conclude

|Ω|nn+1Σ(1+ν,𝐤0)HdA.\displaystyle|\Omega|\leq\frac{n}{n+1}\int_{\Sigma}\frac{(1+\langle\nu,\mathbf{k}_{0}\rangle)}{H}{\rm d}A. (34)

As proved in Theorem 1.1, if equality in (10) holds, then Σ\Sigma is umbilical, and hence spherical. To see that Σ\Sigma must be a θ0\vec{\theta}_{0}-capillary spherical cap, we need a different argument. As equalities hold throughout the argument, we have

|Ω|=|ζ(Z)|=ζ(Z)0(ζ1(y))dy.\displaystyle|\Omega|=|\zeta(Z)|=\int_{\zeta(Z)}\mathcal{H}^{0}(\zeta^{-1}(y)){\rm d}y. (35)

Moreover, (14) implies: for any xΣPix\in\partial\Sigma\cap P_{i}, there holds

(ν(x)+𝐤0),N¯i=cosθi(x)cosθ0i0.\displaystyle\left<-(\nu(x)+\mathbf{k}_{0}),\bar{N}_{i}\right>=\cos\theta^{i}(x)-\cos\theta^{i}_{0}\geq 0. (36)

Recall that N¯i\bar{N}_{i} is the outwards pointing unit normal of PiP_{i}, and we have already showed in the previous proof that Ωζ(Z)\Omega\subset\zeta(Z). Thus, if θi(x)<θ0i\theta^{i}(x)<\theta^{i}_{0} strictly at some xΣPix\in\Sigma\cap P_{i}, then it must be that |Ω|<|ζ(Z)||\Omega|<|\zeta(Z)|, which contradicts to (35). In other words, for any xΣPix\in\partial\Sigma\cap P_{i}, we must have θi(x)=θ0i\theta^{i}(x)=\theta^{i}_{0}, this shows that Σ\Sigma must be a θ0\vec{\theta}_{0}-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 {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0} to test the surjectivity of ζ\zeta, i.e., Ωζ(Z)\Omega\subset\zeta(Z). One subtle point we have to be concerned with is that the first touching of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with Σ\Sigma might occur at ΣP1P2\Sigma\cap P_{1}\cap P_{2}. 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., n+1=3n+1=3, in which case P1P2P_{1}\cap P_{2} is a line.

In the following we use νBr(x)\nu_{B_{r}}(x) to denote the outward unit normal of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}), TxΣT_{x}\Sigma to denote the tangent plane of Σ\Sigma at xx, and ll to denote a unit vector generating the line P1P2P_{1}\cap P_{2}. Recall that N¯i\bar{N}_{i} is the outward unit normal of PiP_{i}, i=1,2.i=1,2.

Case 1. N¯iνBr(x)\bar{N}_{i}\parallel\nu_{B_{r}}(x) for some ii.

Without loss of generality, we assume N2¯\bar{N_{2}} is parallel to νBr(x)\nu_{B_{r}}(x). Hence the sphere Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) touches the plane P2P_{2} only at the point xx. Since N1¯\bar{N_{1}} and N2¯\bar{N_{2}} are not parallel, then N1¯\bar{N_{1}} is not parallel to νBr\nu_{B_{r}}, thus the intersection of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with P1P_{1} must be a circle. Since Sry(y+ry𝐤0)P2={x}S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{2}=\{x\}, we know that Sry(y+ry𝐤0)P1S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{1} touches P1P2P_{1}\cap P_{2} only at xx. Hence ll is the tangential vector of Sry(y+ry𝐤0)P1S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{1} at xx. Since xx is the first touching point of Bry(y+ry𝐤0)B_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with Σ\Sigma from the interior, we see that ll is also the tangential vector of Σ1=ΣP1\partial\Sigma^{1}=\Sigma\cap P_{1} at xx (see Figure 4).

Refer to caption
Figure 4.

It follows that Sry(y+ry𝐤0)P1S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{1} is tangent to Σ1\partial\Sigma^{1} at xx. We are now in the same situation as in Case 2, Theorem 1.1 or Theorem 1.4. Hence the contact angle of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with P1P_{1} at xx must be smaller than θ1(x)\theta^{1}(x), which is a contradiction, since the contact angle of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with P1P_{1} is θ01θ1(x)\theta^{1}_{0}\geq\theta^{1}(x) by assumption.

Case 2. Neither N¯i\bar{N}_{i} is parallel to νBr(x)\nu_{B_{r}}(x), i.e., νBr(x)N¯i0\nu_{B_{r}}(x)\wedge\bar{N}_{i}\neq 0 for i=1,2i=1,2.

In this case, Sry(y+ry𝐤0)Pi(i=1,2)S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{i}(i=1,2) must be circles.

Case 2.1. P1P2TxΣP_{1}\cap P_{2}\subset T_{x}\Sigma.

Since P1P_{1} is not parallel to P2P_{2}, there exists exactly one of Pi(i=1,2)P_{i}(i=1,2), say P1P_{1}, which does not coincide with TxΣT_{x}\Sigma. Then Σ\Sigma is not tangent to P1P_{1} at xx and Σ1+ΣP1\partial\Sigma^{1}+\Sigma\cap P_{1} is a curve near xx. Since lTxΣl\in T_{x}\Sigma and lN¯1l\perp\bar{N}_{1}, we see that ll is the tangential vector of Σ1\partial\Sigma^{1}. Since xx is the first touching point of Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) with Σ\Sigma, ll is also the tangential vector of Sry(y+ry𝐤0)P1S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{1}. Hence Sry(y+ry𝐤0)P1S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{1} is tangent to ΣP1\Sigma\cap P_{1} at xx. 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. P1P2TxΣP_{1}\cap P_{2}\not\subset T_{x}\Sigma.

In this case, N¯i\bar{N}_{i}, i=1,2i=1,2, cannot be parallel to ν(x)\nu(x). For simplicity of notation, in the following we use ν,νBr,θi\nu,\nu_{B_{r}},\theta^{i} to indicate ν(x),νBr(x),θi(x)\nu(x),\nu_{B_{r}}(x),\theta^{i}(x), respectively, and we adopt the notation

Σi=Γi=ΣPi,Bri=Sry(y+ry𝐤0)Pi.\partial\Sigma^{i}=\Gamma_{i}=\Sigma\cap P_{i},\quad\partial B_{r}^{i}=S_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{i}.

Let TΣiT_{\partial\Sigma^{i}} be the unit tangent vector of Σi\partial\Sigma^{i} at xx such that TΣi,Nj¯>0\langle T_{\partial\Sigma^{i}},\bar{N_{j}}\rangle>0 for iji\not=j and TBriT_{\partial B_{r}^{i}} be the unit tangent vector of Bri\partial B_{r}^{i} at xx such that TBri,Nj¯>0\langle T_{\partial B_{r}^{i}},\bar{N_{j}}\rangle>0 for iji\not=j. Since ν\nu and Ni¯\bar{N_{i}} are perpendicular to TΣiT_{\partial\Sigma^{i}}, we know that TΣiT_{\partial\Sigma^{i}} is parallel to νNi¯\nu\wedge\bar{N_{i}}. Similarly, since νBr\nu_{B_{r}} and Ni¯\bar{N_{i}} are perpendicular to TBriT_{\partial B_{r}^{i}}, we have that TBriT_{\partial B_{r}^{i}} is parallel to νBrNi¯\nu_{B_{r}}\wedge\bar{N_{i}}.

Without loss of generality, we may assume that the origin 0Ω0\in\partial\Omega and x0x\neq 0. Let ll be the unit tangent vector x|x|\frac{x}{|x|}. With such choice of ll, we claim that l,ν>0\langle l,\nu\rangle>0 and l,νBr>0\langle l,\nu_{B_{r}}\rangle>0. Indeed, recall that by capillarity, ν\nu is expressed as

ν=cosθiN¯i+sinθiν¯i,i=1,2.\displaystyle\nu=-\cos\theta^{i}\bar{N}_{i}+\sin\theta^{i}\bar{\nu}_{i},\quad i=1,2.

Here ν¯i\bar{\nu}_{i} is the outward unit normal vector of Σi\partial\Sigma^{i} in PiP_{i} at xx. Since Σi\partial\Sigma^{i} lies on the same side of the line P1P2P_{1}\cap P_{2} in PiP_{i}, it yields that l,ν¯i>0\langle l,\bar{\nu}_{i}\rangle>0, thus we get

l,ν=cosθil,Ni¯+sinθil,νi¯=sinθil,νi¯>0.\displaystyle\langle l,\nu\rangle=-\cos\theta^{i}\langle l,\bar{N_{i}}\rangle+\sin\theta^{i}\langle l,\bar{\nu_{i}}\rangle=\sin\theta^{i}\langle l,\bar{\nu_{i}}\rangle>0.

We will show that l,νBr>0\langle l,\nu_{B_{r}}\rangle>0. Since Sry(y+ry𝐤0)S_{r_{y}}(y+r_{y}\mathbf{k}_{0}) touches Σ\Sigma from the interior at xx, we have Bry(y+ry𝐤0)PiΩ¯PiB_{r_{y}}(y+r_{y}\mathbf{k}_{0})\cap P_{i}\subset\bar{\Omega}\cap P_{i} for i=1,2i=1,2. Combining with 0Ω¯P1P20\in\bar{\Omega}\cap P_{1}\cap P_{2}, it follows that ll is pointing outward of BrB_{r} at xx, see Figure 5, thus we obtain l,νBr>0\langle l,\nu_{B_{r}}\rangle>0.

Refer to caption
Figure 5.

Note that N1¯\bar{N_{1}} and N2¯\bar{N_{2}} are perpendicular to ll. Without loss of generality, we may assume

l=N1¯N2¯|N1¯N2¯|.l=-\frac{\bar{N_{1}}\wedge\bar{N_{2}}}{|\bar{N_{1}}\wedge\bar{N_{2}}|}.

A direct computation then yields

νN1¯,N2¯=N1¯N2¯,ν=|N1¯N2¯|l,ν<0,\displaystyle\langle\nu\wedge\bar{N_{1}},\bar{N_{2}}\rangle=\langle\bar{N_{1}}\wedge\bar{N_{2}},\nu\rangle=-|\bar{N_{1}}\wedge\bar{N_{2}}|\langle l,\nu\rangle<0,

which implies the following fact: the two vectors TΣ1T_{\partial\Sigma^{1}} and νN1¯\nu\wedge\bar{N_{1}} are in the opposite direction. Therefore,

TΣ1=νN1¯|νN1¯|.\displaystyle T_{\partial\Sigma^{1}}=-\frac{\nu\wedge\bar{N_{1}}}{|\nu\wedge\bar{N_{1}}|}.

By a similar argument, we obtain

TΣ2=νN2¯|νN2¯|,TBr1=νBrN1¯|νBrN1¯|,TBr2=νBrN2¯|νBrN2¯|.\displaystyle T_{\partial\Sigma^{2}}=\frac{\nu\wedge\bar{N_{2}}}{|\nu\wedge\bar{N_{2}}|},\quad T_{\partial B_{r}^{1}}=-\frac{\nu_{B_{r}}\wedge\bar{N_{1}}}{|\nu_{B_{r}}\wedge\bar{N_{1}}|},\quad T_{\partial B_{r}^{2}}=\frac{\nu_{B_{r}}\wedge\bar{N_{2}}}{|\nu_{B_{r}}\wedge\bar{N_{2}}|}.

Thanks to the fact that xx is the first touching point, we must have (see Figure 6)

TBri,lTΣi,l,i=1,2.\displaystyle\langle T_{\partial B_{r}^{i}},l\rangle\geq\langle T_{\partial\Sigma^{i}},l\rangle,\quad i=1,2. (37)
Refer to caption
Figure 6.

Let ηi(0,π)\eta^{i}\in(0,\pi) be such that νBr,Ni¯=cosηi\langle\nu_{B_{r}},\bar{N_{i}}\rangle=-\cos\eta^{i} for each i=1,2i=1,2. We can carry out the following computations.

TBr1,l=\displaystyle\langle T_{\partial B_{r}^{1}},l\rangle= νBrN1¯|νBrN1¯|,N1¯N2¯|N1¯N2¯|\displaystyle\langle-\frac{\nu_{B_{r}}\wedge\bar{N_{1}}}{|\nu_{B_{r}}\wedge\bar{N_{1}}|},-\frac{\bar{N_{1}}\wedge\bar{N_{2}}}{|\bar{N_{1}}\wedge\bar{N_{2}}|}\rangle
=\displaystyle= νBr,N2¯νBr,N1¯N1¯,N2¯|νBrN1¯||N1¯N2¯|\displaystyle-\frac{\langle\nu_{B_{r}},\bar{N_{2}}\rangle-\langle\nu_{B_{r}},\bar{N_{1}}\rangle\langle\bar{N_{1}},\bar{N_{2}}\rangle}{|\nu_{B_{r}}\wedge\bar{N_{1}}|\cdot|\bar{N_{1}}\wedge\bar{N_{2}}|}
=\displaystyle= cosη2+cosη1cosαsinη1sinα.\displaystyle\frac{\cos\eta^{2}+\cos\eta^{1}\cos\alpha}{\sin\eta^{1}\sin\alpha}.

Similarly,

TBr2,l=cosη1+cosη2cosαsinη2sinα,\displaystyle\langle T_{\partial B_{r}^{2}},l\rangle=\frac{\cos\eta^{1}+\cos\eta^{2}\cos\alpha}{\sin\eta^{2}\sin\alpha},
TΣ1,l=cosθ2+cosθ1cosαsinθ1sinα,\displaystyle\langle T_{\partial\Sigma^{1}},l\rangle=\frac{\cos\theta^{2}+\cos\theta^{1}\cos\alpha}{\sin\theta^{1}\sin\alpha},
TΣ2,l=cosθ1+cosθ2cosαsinθ2sinα.\displaystyle\langle T_{\partial\Sigma^{2}},l\rangle=\frac{\cos\theta^{1}+\cos\theta^{2}\cos\alpha}{\sin\theta^{2}\sin\alpha}.

Plugging into (37), we thus obtain

cosη2+cosη1cosαsinη1sinαcosθ2+cosθ1cosαsinθ1sinα,\displaystyle\frac{\cos\eta^{2}+\cos\eta^{1}\cos\alpha}{\sin\eta^{1}\sin\alpha}\geq\frac{\cos\theta^{2}+\cos\theta^{1}\cos\alpha}{\sin\theta^{1}\sin\alpha}, (38)
cosη1+cosη2cosαsinη2sinαcosθ1+cosθ2cosαsinθ2sinα.\displaystyle\frac{\cos\eta^{1}+\cos\eta^{2}\cos\alpha}{\sin\eta^{2}\sin\alpha}\geq\frac{\cos\theta^{1}+\cos\theta^{2}\cos\alpha}{\sin\theta^{2}\sin\alpha}. (39)

A crucial observation is that (38) is equivalent to (see the end of Appendix A)

sinθ1(cosη2\displaystyle\sin\theta^{1}\big{(}\cos\eta^{2} cos(θ2θ1+η1))\displaystyle-\cos(\theta^{2}-\theta^{1}+\eta^{1})\big{)}
+\displaystyle+ (cos(θ2θ1)+cosα)sin(θ1η1)0.\displaystyle\big{(}\cos(\theta^{2}-\theta^{1})+\cos\alpha\big{)}\sin(\theta^{1}-\eta^{1})\geq 0. (40)

On the one hand, we have

x,N1¯=\displaystyle\langle x,\bar{N_{1}}\rangle= x(y+ry𝐤0),N1¯+y+ry𝐤0,N1¯\displaystyle\langle x-(y+r_{y}\mathbf{k}_{0}),\bar{N_{1}}\rangle+\langle y+r_{y}\mathbf{k}_{0},\bar{N_{1}}\rangle
=\displaystyle= ryνBr,N1¯+y,N1¯+ry𝐤0,N1¯.\displaystyle r_{y}\langle\nu_{B_{r}},\bar{N_{1}}\rangle+\langle y,\bar{N_{1}}\rangle+r_{y}\langle\mathbf{k}_{0},\bar{N_{1}}\rangle. (41)

Since xP1x\in P_{1} and y𝐖̊y\in\mathring{\mathbf{W}}, we have: x,N1¯=0\langle x,\bar{N_{1}}\rangle=0 and y,N1¯<0\langle y,\bar{N_{1}}\rangle<0. It follows from (4) that

νBr,N1¯>𝐤0,N1¯=cosθ01.\displaystyle\langle\nu_{B_{r}},\bar{N_{1}}\rangle>-\langle\mathbf{k}_{0},\bar{N_{1}}\rangle=-\cos\theta_{0}^{1}.

By our angle assumption θ1θ01\theta^{1}\leq\theta_{0}^{1}, we get

cosη1=νBr,N1¯>cosθ01cosθ1,\displaystyle-\cos\eta^{1}=\langle\nu_{B_{r}},\bar{N_{1}}\rangle>-\cos\theta_{0}^{1}\geq-\cos\theta^{1},

which implies

η1>θ1.\displaystyle\eta^{1}>\theta^{1}. (42)

On the other hand, since P1P2TxΣP_{1}\cap P_{2}\not\subset T_{x}\Sigma, by Lemma A.2, we see that |𝐤(x)|<1|\mathbf{k}(x)|<1, where 𝐤(x):=i=12ci(x)N¯i\mathbf{k}(x):=\sum\limits_{i=1}^{2}c_{i}(x)\bar{N}_{i}, satisfying 𝐤(x),N¯i=cosθi\langle\mathbf{k}(x),\bar{N}_{i}\rangle=\cos\theta^{i}. In view of Lemma A.1, it implies that

cos(θ2θ1)+cosα>0.\displaystyle\cos(\theta^{2}-\theta^{1})+\cos\alpha>0. (43)

From (42) and (43), we know that

(cos(θ2θ1)+cosα)sin(θ1η1)<0.\big{(}\cos(\theta^{2}-\theta^{1})+\cos\alpha\big{)}\sin(\theta^{1}-\eta^{1})<0.

Combining with (4), we find

cosη2cos(θ2θ1+η1)>0.\displaystyle\cos\eta^{2}-\cos(\theta^{2}-\theta^{1}+\eta^{1})>0. (44)

By a similar argument we obtain from (39) that

η2>θ2,\displaystyle\eta^{2}>\theta^{2}, (45)

and

cosη1cos(θ1θ2+η2)>0.\displaystyle\cos\eta^{1}-\cos(\theta^{1}-\theta^{2}+\eta^{2})>0. (46)

We may assume θ2θ1\theta^{2}\geq\theta^{1} without loss of generality, then it follows from (45) that

0<θ1<θ1θ2+η2η2<π.\displaystyle 0<\theta^{1}<\theta^{1}-\theta^{2}+\eta^{2}\leq\eta^{2}<\pi.

Therefore, back to (44), we deduce that

η1<θ1θ2+η2,\displaystyle\eta^{1}<\theta^{1}-\theta^{2}+\eta^{2}, (47)

In the meanwhile, from (46), we deduce

η2<θ2θ1+η1.\displaystyle\eta^{2}<\theta^{2}-\theta^{1}+\eta^{1}. (48)

Apparently, (47) and (48) lead to a contradiction.

In conclusion, we have showed that the first touching point cannot occur at any xΣP1P2x\in\Sigma\cap P_{1}\cap P_{2}. The rest of the proof follows similarly from Theorem 1.4. ∎

Remark 4.1.

We make a remark for the case of higher dimensions n+1\mathbb{R}^{n+1}. In this case, P1P2P_{1}\cap P_{2} is a (n1)(n-1)-plane and ΣP1P2\partial\Sigma\cap P_{1}\cap P_{2} is a closed (n2)(n-2)-submanifold, playing the role as (ΣP1)\partial(\partial\Sigma\cap P_{1}) or (ΣP2)\partial(\partial\Sigma\cap P_{2}). By letting ll be the unique unit vector in P1P2P_{1}\cap P_{2} which is orthogonal to ΣP1P2\partial\Sigma\cap P_{1}\cap P_{2}, and TΣiT_{\partial\Sigma^{i}} be the tangent vector of ΣPi\partial\Sigma\cap P_{i} at xx which is orthogonal to ΣP1P2\partial\Sigma\cap P_{1}\cap P_{2}, and TBriT_{\partial B_{r}^{i}} be the tangent vector of BrPi\partial B_{r}\cap P_{i} at xx which is orthogonal to ΣP1P2=BrP1P2\partial\Sigma\cap P_{1}\cap P_{2}=\partial B_{r}\cap P_{1}\cap P_{2} ( because xx is the first touching point), we may reduce the problem to the 33-dimensional case. In other words, all the vectors at xx we considered in the proof lie in the orthogonal 33-subspace of Tx(ΣP1P2)T_{x}(\partial\Sigma\cap P_{1}\cap P_{2}).

5. Alexandrov-Type Theorem

It is well-known that there exists at least one elliptic point for a closed embedded hypersurface in n+1\mathbb{R}^{n+1}. 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 𝐖\mathbf{W} be a generalized wedge and yi=1LPiy\in\bigcap\limits_{i=1}^{L}P_{i}. If |𝐤0|<1|\mathbf{k}_{0}|<1, then for any r>0r>0, the sphere Sr(y+r𝐤0)S_{r}(y+r\mathbf{k}_{0}) intersects PiP_{i} at angle θ0i\theta_{0}^{i}. In particular, in +n+1¯\overline{\mathbb{R}^{n+1}_{+}}, the sphere Sr(yrcosθ0En+1)S_{r}(y-r\cos\theta_{0}E_{n+1}) intersects +n+1\partial\mathbb{R}^{n+1}_{+} at angle θ0\theta_{0}.

Proof.

First, if |𝐤0|<1|\mathbf{k}_{0}|<1, it is easy to see that for any r>0r>0, Sr(y+r𝐤0)S_{r}(y+r\mathbf{k}_{0}) intersects P̊i\mathring{P}_{i}. The outward unit normal to Sr(y+r𝐤0)S_{r}(y+r\mathbf{k}_{0}) at xP̊ix\in\mathring{P}_{i} is given by

νBr(x)=x(y+r𝐤0)r.\displaystyle\nu_{B_{r}}(x)=\frac{x-(y+r\mathbf{k}_{0})}{r}.

Since xyPix-y\in P_{i}, we see that

νBr(x),N¯i=𝐤0,N¯i=cosθ0i.\langle\nu_{B_{r}}(x),\bar{N}_{i}\rangle=-\langle\mathbf{k}_{0},\bar{N}_{i}\rangle=-\cos\theta_{0}^{i}.

Proposition 5.2.

Let Σ+n+1¯\Sigma\subset\overline{\mathbb{R}^{n+1}_{+}} be a smooth, compact, embedded θ0\theta_{0}-capillary hypersurface. Then there exists at least one elliptic point on Σ\Sigma.

Proof.

Let Ω\Omega be the enclosed region of Σ\Sigma and +n+1\partial\mathbb{R}^{n+1}_{+}, T=Ω+n+1T=\partial\Omega\cap\partial\mathbb{R}^{n+1}_{+}, we fix a point yT̊y\in\mathring{T}. Consider the family of the open balls Br(yrcosθ0En+1)B_{r}(y-r\cos\theta_{0}E_{n+1}). By Lemma 5.1, Sr(yrcosθ0En+1)=Br(yrcosθ0En+1)S_{r}(y-r\cos\theta_{0}E_{n+1})=\partial B_{r}(y-r\cos\theta_{0}E_{n+1}) intersects with +n+1\partial\mathbb{R}^{n+1}_{+} at the angle θ0\theta_{0}. Since Σ\Sigma is compact, for rr large enough, ΣBr(yrcosθ0En+1)\Sigma\subset B_{r}(y-r\cos\theta_{0}E_{n+1}). Hence we can find the smallest rr, say r0>0r_{0}>0, such that Sr0(yr0cosθ0En+1)S_{r_{0}}(y-r_{0}\cos\theta_{0}E_{n+1}) touches Σ\Sigma at a first time at some xΣx\in\Sigma. For simplicity, we abbreviate Sr0(yr0cosθ0En+1)S_{r_{0}}(y-r_{0}\cos\theta_{0}E_{n+1}) by Sr0S_{r_{0}}.

If xΣ̊x\in\mathring{\Sigma}, then Σ\Sigma and Sr0S_{r_{0}} are tangent at xx. If xΣx\in\partial\Sigma, from the fact that both Σ\Sigma and Sr0S_{r_{0}} intersect with +n+1\partial\mathbb{R}^{n+1}_{+} at the angle θ0\theta_{0}, we conclude again that Σ\Sigma and Sr0S_{r_{0}} are tangent at xx.

In both cases, we have that the principal curvatures of Σ\Sigma at xx are bigger than or equal to 1/r0>01/r_{0}>0, which implies that xx is an elliptic point. ∎

Proposition 5.3.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge. Let Σ𝐖\Sigma\subset\mathbf{W} be a smooth, compact, embedded θ0\vec{\theta}_{0}-capillary surface. If ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset, then there exists at least one elliptic point on Σ\Sigma.

Proof.

In view of Remark 4.1, we need only consider 33-dimensional case.

Since the corners Γi\Gamma_{i} (i=1,2i=1,2) are smooth co-dimension two submanifolds in +n+1\partial\mathbb{R}^{n+1}_{+}, Σ\partial\Sigma is embedded in n+1\mathbb{R}^{n+1} (see Section 2), when ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset, we must have that ΩP1P2\partial\Omega\cap P_{1}\cap P_{2} is also a co-dimension two submanifold in n+1\mathbb{R}^{n+1}. This means it has non-trivial interior relative to the topology of P1P2P_{1}\cap P_{2}, therefore we are free to choose any yy in the interior of ΩP1P2\partial\Omega\cap P_{1}\cap P_{2} as the initial center of the sphere foliation {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0}. Thanks to Lemma A.4, any such foliation must have a first touching point with Σ\Sigma from outside as rr decreases from \infty.

We consider the sphere Sr0(y+r0𝐤0)S_{r_{0}}(y+r_{0}\mathbf{k}_{0}) which touches Σ\Sigma for the first time at some xΣx\in\Sigma. Following the argument in LABEL:Prop-ellipticpoint1, we see that, if xΣ̊x\in\mathring{\Sigma} or xΣ(P1P2)x\in\partial\Sigma\setminus(P_{1}\cap P_{2}), then xx must be an elliptic point of Σ\Sigma. Therefore, it remains to consider the case xΣ(P1P2)x\in\partial\Sigma\cap(P_{1}\cap P_{2}).

Since x,yP1P2x,y\in P_{1}\cap P_{2}, we find

νBr(x),N¯1=cosθ01=ν(x),N1¯,\displaystyle\left<\nu_{B_{r}}(x),-\bar{N}_{1}\right>=\cos\theta_{0}^{1}=\langle\nu(x),-\bar{N_{1}}\rangle,
νBr(x),N¯2=cosθ02=ν(x),N2¯.\displaystyle\left<\nu_{B_{r}}(x),-\bar{N}_{2}\right>=\cos\theta_{0}^{2}=\langle\nu(x),-\bar{N_{2}}\rangle.

This implies that either (a) ν(x)=νBr(x)\nu(x)=\nu_{B_{r}}(x), or (b) ν(x)=νBr(x)+al\nu(x)=\nu_{B_{r}}(x)+al, where ll is a unit vector perpendicular to both N1¯\bar{N_{1}} and N2¯\bar{N_{2}}, and aa is a non-zero constant.

We claim that the situation (b) does not occur. Otherwise, by the fact that ν(x)\nu(x) and νBr(x)\nu_{B_{r}}(x) are unit vectors, we can deduce that a=2νBr(x),l0a=-2\langle\nu_{B_{r}(x)},l\rangle\neq 0. Thus we have

ν(x),l=νBr(x)+al,l=νBr(x),l0.\displaystyle\langle\nu(x),l\rangle=\langle\nu_{B_{r}}(x)+al,l\rangle=-\langle\nu_{B_{r}}(x),l\rangle\neq 0. (49)

Note that xyx\neq y and xyΩ¯Br0(y+r0𝐤0)x-y\in\overline{\Omega}\cap B_{r_{0}}(y+r_{0}\mathbf{k}_{0}), by the fact that ν(x)\nu(x) and νBr(x)\nu_{B_{r}}(x) are outward normal vectors, we have

ν(x),xy0,νBr(x),xy0.\langle\nu(x),x-y\rangle\geq 0,\quad\langle\nu_{B_{r}}(x),x-y\rangle\geq 0.

Since x,yP1P2x,y\in P_{1}\cap P_{2}, ll is parallel to xyx-y. It follows that ν(x),l\langle\nu(x),l\rangle and νBr(x),l\langle\nu_{B_{r}}(x),l\rangle have the same sign, which contradicts (49) and concludes the claim.

From the claim, we see that only the situation (a) happens. This means, SrS_{r} and Σ\Sigma are tangent at xx. Thus the principal curvatures of Σ\Sigma at xx are bigger than or equal to 1/r0>01/r_{0}>0, which implies xx is an elliptic point. ∎

In view of the proof of LABEL:Prop-ellipticpoint2, the only place where we used the condition ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset is that we have to use it in Lemma A.4, as |𝐤0|=1|\mathbf{k}_{0}|=1, to conclude that the sphere foliation {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0} must touch Σ\Sigma. In this regard, we can remove the extra assumption ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset by strengthening the angle condition to be |𝐤0|<1|\mathbf{k}_{0}|<1. Indeed, we have the following

Proposition 5.4.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge. Let Σ𝐖\Sigma\subset\mathbf{W} be a smooth, compact, embedded θ0\vec{\theta}_{0}-capillary hypersurface with |𝐤0|<1|\mathbf{k}_{0}|<1. Then there exists at least one elliptic point on Σ\Sigma.

If the hypersurfaces are CMC hypersurfaces, we have

Proposition 5.5.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge. Let Σ𝐖\Sigma\subset\mathbf{W} be a smooth, compact, embedded θ0\vec{\theta}_{0}-capillary hypersurface. If Σ\Sigma is of constant mean curvature HH, then H>0H>0, provided παθ01+θ02.\pi-\alpha\leq\theta_{0}^{1}+\theta_{0}^{2}.

Proof.

We remark that the free boundary case, that is θ0=(π2,π2)\vec{\theta}_{0}=(\frac{\pi}{2},\frac{\pi}{2}) has been proved by Lopez [Lopez14]. We consider here the general case.

To begin, we take p0Σp_{0}\in\partial\Sigma to be the point of maximal distance of Σ\partial\Sigma from the edge P1P2P_{1}\cap P_{2}. Assume without loss of generality that p0P2p_{0}\in P_{2}. Let 𝒫\mathcal{P} be the family of planes parallel to the edge of 𝐖\mathbf{W} and having contact angle θ02\theta_{0}^{2} with P2P_{2}. Starting from one of such a plane near infinite and moving it among this family until the first time that one of plane PP in 𝒫\mathcal{P} touches Σ\Sigma. By definition of p0p_{0} and PP, it is only possible that the first touching occurs at certain interior point of Σ\Sigma, at p0P2p_{0}\in P_{2} or at some p1ΣP1p_{1}\in\Sigma\cap P_{1}.

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 H>0H>0. For the last case, we shall use our angle assumption παθ1+θ2\pi-\alpha\leq\theta_{1}+\theta_{2}.

Indeed, if πα<θ1+θ2\pi-\alpha<\theta_{1}+\theta_{2}, then the first touching point must not occur at any points of ΣP1\Sigma\cap P_{1}: since PP is a plane having contact angle θ02\theta_{0}^{2} with P2P_{2}, we know that the contact angle of Σ\Sigma with P1P_{1}, say θP1\theta_{P}^{1}, is παθ02\pi-\alpha-\theta_{0}^{2}. By the angle assumption, we find

θP1=παθ02<θ01,\displaystyle\theta_{P}^{1}=\pi-\alpha-\theta_{0}^{2}<\theta_{0}^{1}, (50)

which is not possible if PP touches ΣP1\Sigma\cap P_{1} from outside for the first time.

If πα=θ01+θ02\pi-\alpha=\theta_{0}^{1}+\theta_{0}^{2}, then θP1=θ01\theta_{P}^{1}=\theta_{0}^{1} from the above discussion, and we can use the Hopf’s boundary point lemma again to find that H>0H>0. 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 |𝐤0|1|\mathbf{k}_{0}|\leq 1 ensures the validity of the inequality

1+ν,𝐤001+\left<\nu,\mathbf{k}_{0}\right>\geq 0

pointwisely on Σ\Sigma. In particular, as L=1L=1, we have

1cosθ0ν,En+10\displaystyle 1-\cos\theta_{0}\left<\nu,E_{n+1}\right>\geq 0

along Σ\Sigma.

On one hand, by virtue of LABEL:Prop-ellipticpoint2 and Gärding’s argument [Garding59] (see also [Ros87, Section 3]), we know that HjH_{j} are positive, for jrj\leq r. Applying Theorem 1.5 and using the Maclaurin inequality

H1Hr1/r,H_{1}\geq H_{r}^{1/r},

we find

(n+1)Hr1/r|Ω|Hr1/rΣ1+ν,𝐤0H/ndAΣ(1+ν,𝐤0)dA,\displaystyle(n+1)H_{r}^{1/r}|\Omega|\leq H_{r}^{1/r}\int_{\Sigma}\frac{1+\left<\nu,\mathbf{k}_{0}\right>}{H/n}{\rm d}A\leq\int_{\Sigma}\left(1+\left<\nu,\mathbf{k}_{0}\right>\right){\rm d}A, (51)

and equality holds if and only if Σ\Sigma is a θ0{\theta}_{0}-capillary spherical cap.

On the other hand, using the Minkowski formula (23) and the Maclaurin inequality again, we have

0=\displaystyle 0= Σ(1+ν,𝐤0)Hr1Hrx,νdA\displaystyle\int_{\Sigma}\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)H_{r-1}-H_{r}\left<x,\nu\right>{\rm d}A
\displaystyle\geq Σ(1+ν,𝐤0)Hrr1rHrx,νdA\displaystyle\int_{\Sigma}\left(1+\left<\nu,\mathbf{k}_{0}\right>\right)H_{r}^{\frac{r-1}{r}}-H_{r}\left<x,\nu\right>{\rm d}A
=\displaystyle= Hrr1rΣ1+ν,𝐤0Hr1/rx,νdA\displaystyle H_{r}^{\frac{r-1}{r}}\int_{\Sigma}1+\left<\nu,\mathbf{k}_{0}\right>-H_{r}^{1/r}\left<x,\nu\right>{\rm d}A
=\displaystyle= Hrr1r{Σ1+ν,𝐤0dA(n+1)Hr1/r|Ω|},\displaystyle H_{r}^{\frac{r-1}{r}}\left\{\int_{\Sigma}1+\left<\nu,\mathbf{k}_{0}\right>{\rm d}A-(n+1)H_{r}^{1/r}|\Omega|\right\},

where in the last equality we have used (32). Thus equality in (51) holds, and hence Σ\Sigma is a θ0\vec{\theta}_{0}-capillary spherical cap. This completes the proof. ∎

Proof of Theorem 1.8.

For the CMC case, we notice that our condition (8) implies παθ01+θ02\pi-\alpha\leq\theta_{0}^{1}+\theta_{0}^{2} automatically, thanks to Lemma A.1. Therefore, we can use Proposition 5.5 to see that Σ\Sigma is of positive constant mean curvature.

For the constant HrH_{r} case, since we assume |𝐤0|<1|\mathbf{k}_{0}|<1, we may use Proposition 5.4 to conclude that Σ\Sigma has an elliptic point, from which we see that HrH_{r} 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 Σ\Sigma must be a θ0\vec{\theta}_{0}-spherical cap. However, if this is the case, a simple geometric relation (see Figure 7) then implies that α+(πθ01)+(πθ02)<π\alpha+(\pi-\theta_{0}^{1})+(\pi-\theta_{0}^{2})<\pi, i.e., π+α<θ01+θ02\pi+\alpha<\theta_{0}^{1}+\theta_{0}^{2}, which is incompatible with (11). The proof is complete. ∎

Refer to caption
Figure 7.

Appendix A Miscellaneous Results in Wedge

Lemma A.1.

For the case L=2L=2, namely, 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} is a classical wedge, (8) is equivalent to (11), i.e.,

|π(θ01+θ02)|απ|θ01θ02|.\displaystyle|\pi-(\theta_{0}^{1}+\theta_{0}^{2})|\leq\alpha\leq\pi-|\theta_{0}^{1}-\theta_{0}^{2}|.

Similarly, (8) with strict inequality is equivalent to (11) with strict inequality.

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 α\alpha.) with prescribed angles θ0=(θ01,θ02)\vec{\theta_{0}}=(\theta^{1}_{0},\theta^{2}_{0}).

By definition of 𝐤0\mathbf{k}_{0} and the fact that N¯1,N¯2=cos(πα)=cosα\left<\bar{N}_{1},\bar{N}_{2}\right>=\cos({\pi-\alpha})=-\cos\alpha, we find

{c1c2cosα=cosθ01,c1cosα+c2=cosθ02.\displaystyle\begin{cases}c_{1}-c_{2}\cos\alpha&=\cos\theta^{1}_{0},\\ -c_{1}\cos\alpha+c_{2}&=\cos\theta^{2}_{0}.\end{cases}

A simple computation then yields

(c1c2)=1sin2α(1cosαcosα1)(cosθ01cosθ02),\displaystyle\begin{pmatrix}c_{1}\\ c_{2}\end{pmatrix}=\frac{1}{\sin^{2}\alpha}\begin{pmatrix}1&\cos\alpha\\ \cos\alpha&1\end{pmatrix}\begin{pmatrix}\cos\theta^{1}_{0}\\ \cos\theta^{2}_{0}\end{pmatrix},

and it follows that

|𝐤0|2=\displaystyle|\mathbf{k}_{0}|^{2}= 𝐤0,c1N¯1+c2N¯2=c1cosθ01+c2cosθ02\displaystyle\left<\mathbf{k}_{0},c_{1}\bar{N}_{1}+c_{2}\bar{N}_{2}\right>=c_{1}\cos\theta^{1}_{0}+c_{2}\cos\theta^{2}_{0}
=\displaystyle= (cosθ01cosθ02)(c1c2)\displaystyle\begin{pmatrix}\cos\theta^{1}_{0}&\cos\theta^{2}_{0}\end{pmatrix}\begin{pmatrix}c_{1}\\ c_{2}\end{pmatrix}
=\displaystyle= cos2θ01+cos2θ02+2cosθ01cosθ02cosαsin2α.\displaystyle\frac{\cos^{2}{\theta_{0}^{1}}+\cos^{2}\theta_{0}^{2}+2\cos{\theta_{0}^{1}}\cos{\theta_{0}^{2}}\cos{\alpha}}{\sin^{2}{\alpha}}.

Thus (8) can be rewritten as

cos2θ01+cos2θ02+2cosθ01cosθ02cosαsin2α.\displaystyle\cos^{2}{\theta_{0}^{1}}+\cos^{2}\theta_{0}^{2}+2\cos{\theta_{0}^{1}}\cos{\theta_{0}^{2}}\cos{\alpha}\leq\sin^{2}{\alpha}.

which is

(cosα+cos(θ01+θ02))(cosα+cos(θ01θ02))0.\displaystyle\big{(}\cos\alpha+\cos(\theta_{0}^{1}+\theta_{0}^{2})\big{)}\big{(}\cos\alpha+\cos(\theta_{0}^{1}-\theta_{0}^{2})\big{)}\leq 0.

Thus we see that (8) is equivalent to

|θ01θ02|παθ01+θ02π+α,\displaystyle|\theta_{0}^{1}-\theta_{0}^{2}|\leq\pi-\alpha\leq\theta_{0}^{1}+\theta_{0}^{2}\leq\pi+\alpha,

which is just (11). ∎

The following observation is important for our analysis in Theorem 1.5.

Lemma A.2.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge. If Σ𝐖\Sigma\subset\mathbf{W} is a smooth, compact, embedded θ\vec{\theta}-capillary hypersurface. Let 𝐤:Σn+1\mathbf{k}:\partial\Sigma\to\mathbb{R}^{n+1} be given by 𝐤(x)=i=12ci(x)N¯i\mathbf{k}(x)=\sum\limits_{i=1}^{2}c_{i}(x)\bar{N}_{i} such that 𝐤(x),N¯=cosθi(x).\langle\mathbf{k}(x),\bar{N}\rangle=\cos\theta^{i}(x). If ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset, then we have |𝐤|1|\mathbf{k}|\leq 1 on ΣP1P2\Sigma\cap P_{1}\cap P_{2}. Moreover, for any xΣP1P2x\in\Sigma\cap P_{1}\cap P_{2}, P1P2TxΣP_{1}\cap P_{2}\subset T_{x}\Sigma if and only if |𝐤(x)|=1|\mathbf{k}(x)|=1.

Proof.

In view of Remark 4.1, we need only consider the 33-dimensional case.

Let xΣP1P2x\in\Sigma\cap P_{1}\cap P_{2}, In the following, we compute at xx. We have

N1¯,N2¯=cosα,ν,Ni¯=cosθi,i=1,2.\langle\bar{N_{1}},\bar{N_{2}}\rangle=-\cos\alpha,\quad\langle\nu,\bar{N_{i}}\rangle=-\cos\theta^{i},\quad i=1,2.

Thus

N1¯ν,N2¯ν\displaystyle\langle\bar{N_{1}}\wedge\nu,\bar{N_{2}}\wedge\nu\rangle =N1¯,N2¯N1¯,νN2¯,ν\displaystyle=\langle\bar{N_{1}},\bar{N_{2}}\rangle-\langle\bar{N_{1}},\nu\rangle\langle\bar{N_{2}},\nu\rangle
=cosαcosθ1cosθ2.\displaystyle=-\cos\alpha-\cos\theta^{1}\cos\theta^{2}.

Since |N1¯ν,N2¯ν|sinθ1sinθ2|\langle\bar{N_{1}}\wedge\nu,\bar{N_{2}}\wedge\nu\rangle|\leq\sin\theta^{1}\sin\theta^{2}, we deduce

|cosα+cosθ1cosθ2|sinθ1sinθ2,\displaystyle|\cos\alpha+\cos\theta^{1}\cos\theta^{2}|\leq\sin\theta^{1}\sin\theta^{2},

which implies |𝐤|1|\mathbf{k}|\leq 1.

If P1P2TxΣP_{1}\cap P_{2}\subset T_{x}\Sigma, we have |N1¯ν,N2¯ν|=sinθ1sinθ2|\langle\bar{N_{1}}\wedge\nu,\bar{N_{2}}\wedge\nu\rangle|=\sin\theta^{1}\sin\theta^{2}. Then we get

|cosα+cosθ1cosθ2|=sinθ1sinθ2,\displaystyle|\cos\alpha+\cos\theta^{1}\cos\theta^{2}|=\sin\theta^{1}\sin\theta^{2},

which is equivalent to |𝐤(x)|=1|\mathbf{k}(x)|=1. ∎

We point out that, in the discussion above, if Σ\Sigma intersects P1P_{1} and P2P_{2} transversally at xP1P2x\in P_{1}\cap P_{2}, then Lemma A.2 is included in [Li21, Lemma 2.5].

Lemma A.3.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge. If Σ𝐖\Sigma\subset\mathbf{W} is a smooth, compact, embedded θ0\vec{\theta}_{0}-capillary with |𝐤0|1|\mathbf{k}_{0}|\leq 1, then for any yΩy\in\Omega, the family of spheres {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0} must touch Σ\Sigma.

Proof.

The case |𝐤0|<1|\mathbf{k}_{0}|<1 follows trivially, since {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0} foliates the whole n+1\mathbb{R}^{n+1}. As for |𝐤0|=1|\mathbf{k}_{0}|=1, we proceed by the following observation.

Observation. Since ySr(y+r𝐤0)y\in S_{r}(y+r\mathbf{k}_{0}) for any r0r\geq 0, with νSr(y)=𝐤0\nu_{S_{r}}(y)=-\mathbf{k}_{0}. Moreover,

Br(y+r𝐤0)Hy:={zn+1:zy,𝐤0>0}as r.\displaystyle B_{r}(y+r\mathbf{k}_{0})\rightarrow H_{y}^{-}:=\{z\in\mathbb{R}^{n+1}:\left<z-y,\mathbf{k}_{0}\right>>0\}\quad\text{as }r\rightarrow\infty. (52)

In other words, the family of spheres Sr(y+r𝐤0)S_{r}(y+r\mathbf{k}_{0}) foliates the half-space HyH^{-}_{y}.

We claim that for any yΩy\in\Omega, there holds HyΣH_{y}^{-}\cap\Sigma\neq\emptyset. To see this, we consider the following situations separately.

Case 1. ΣP1P2=\Sigma\cap P_{1}\cap P_{2}=\emptyset.

By definition of Σ\Sigma, we see that Ω=ΣT1T2\partial\Omega=\Sigma\cup T_{1}\cup T_{2} with T1T_{1}, T2T_{2} away from the edge P1P2P_{1}\cap P_{2} (see Figure 7 for illustration). Therefore for any 𝐤0𝕊n\mathbf{k}_{0}\in\mathbb{S}^{n}, the open half-space determined by y,𝐤0y,\mathbf{k}_{0} must intersect Σ\Sigma.

Case 2. ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset.

In this case, since ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset, we must have P1P2TxΣP_{1}\cap P_{2}\subset T_{x}\Sigma for any xΣP1P2x\in\Sigma\cap P_{1}\cap P_{2}, according to Lemma A.2 (with θ=θ0\vec{\theta}=\vec{\theta}_{0} chosen therein). However, this contradicts to our assumption on Σ\Sigma, precisely, that Γi\Gamma_{i} (i=1,2)(i=1,2) are smooth co-dimension two submanifolds in n+1\mathbb{R}^{n+1}. This proves the claim and hence completes the proof.

In the proof of LABEL:Prop-ellipticpoint2, we use the family of spheres Sr(y+r𝐤0)S_{r}(y+r\mathbf{k}_{0}) for some yy in the interior of Ω¯P1P2\bar{\Omega}\cap P_{1}\cap P_{2}, the following observation is needed.

Lemma A.4.

Let 𝐖n+1\mathbf{W}\subset\mathbb{R}^{n+1} be a classical wedge. If Σ𝐖\Sigma\subset\mathbf{W} is a smooth, compact, embedded θ0\vec{\theta}_{0}-capillary such that ΣP1P2\Sigma\cap P_{1}\cap P_{2}\neq\emptyset, then for any yy in the interior of ΩP1P2\partial\Omega\cap P_{1}\cap P_{2}, the family of spheres {Sr(y+r𝐤0)}r0\{S_{r}(y+r\mathbf{k}_{0})\}_{r\geq 0} must touch Σ\Sigma.

Proof.

In view of Remark 4.1, we need only consider the 33-dimensional case.

By virtue of Lemma A.2, we have |𝐤0|1|\mathbf{k}_{0}|\leq 1. The case |𝐤0|<1|\mathbf{k}_{0}|<1 follows trivially from Lemma A.3. As for |𝐤0|=1|\mathbf{k}_{0}|=1, by Observation above and the fact that y,𝐤0=0\langle y,\mathbf{k}_{0}\rangle=0 due to yP1P2y\in P_{1}\cap P_{2}, it suffices to show that P1,P2H0P_{1},P_{2}\subset H_{0}^{-}.

Claim. In (7), we have c1,c2<0c_{1},c_{2}<0, provided that |𝐤0|=1|\mathbf{k}_{0}|=1 and ΣP1P2.\Sigma\cap P_{1}\cap P_{2}\neq\emptyset.

If the claim holds, a direct computation yields: for any zP1̊z\in\mathring{P_{1}},

z,𝐤0=z,c1N¯1+c2N¯2=c2z,N¯2.\displaystyle\langle z,\mathbf{k}_{0}\rangle=\langle z,c_{1}\bar{N}_{1}+c_{2}\bar{N}_{2}\rangle=c_{2}\langle z,\bar{N}_{2}\rangle.

Notice that {l,l1,N¯1}\{l,l_{1},\bar{N}_{1}\} forms an orthonormal basis of 3\mathbb{R}^{3}, where ll is a fixed unit vector, parallel to P1P2P_{1}\cap P_{2}, l1l_{1} is the unit inwards pointing conormal of P1\partial P_{1} in P1P_{1}. In this coordinate, since zP1̊z\in\mathring{P_{1}}, it can be expressed as z=a0l+a1l1+0N¯1z=a_{0}l+a_{1}l_{1}+0\bar{N}_{1} with a1>0a_{1}>0. It follows that z,𝐤0=a1c2(sinα)>0\langle z,\mathbf{k}_{0}\rangle=a_{1}c_{2}(-\sin\alpha)>0. Similarly, we have: for any zP2̊z\in\mathring{P_{2}}, there holds z,𝐤0>0\langle z,\mathbf{k}_{0}\rangle>0, which implies that P1,P2H0P_{1},P_{2}\subset H_{0}^{-} and proves the Lemma.

We are thus left to prove the Claim. Indeed, by Lemma A.2, let xΣP1P2x\in\Sigma\cap P_{1}\cap P_{2}, then we have P1P2TxΣP_{1}\cap P_{2}\in T_{x}\Sigma. In view of this, we obtain: ν(x)span{N¯1,N¯2}\nu(x)\in{\rm span}\{\bar{N}_{1},\bar{N}_{2}\}, where ν(x)\nu(x) is the unit outward normal of Σ\Sigma at xx. Due to the contact angle condition, we must have

ν(x),N¯i=cosθ0i,\displaystyle\langle\nu(x),\bar{N}_{i}\rangle=-\cos\theta_{0}^{i}, (53)

comparing with the definition of 𝐤0\mathbf{k}_{0} (7), we thus find:

ν(x)=𝐤0=c1N¯1c2N¯2.\nu(x)=-\mathbf{k}_{0}=-c_{1}\bar{N}_{1}-c_{2}\bar{N}_{2}.

Since ν(x)\nu(x) is the outward unit normal of Σ\Sigma at xx, for any yiΩPi̊,i=1,2y_{i}\in\partial\Omega\cap\mathring{P_{i}},i=1,2, we have

xyi,ν(x)>0.\displaystyle\langle x-y_{i},\nu(x)\rangle>0.

Indeed, it follows from ν(x)=cosθ0iN¯i+sinθ0iν¯i(x)\nu(x)=-\cos\theta_{0}^{i}\bar{N}_{i}+\sin\theta_{0}^{i}\bar{\nu}_{i}(x) and the fact ν¯i(x)\bar{\nu}_{i}(x) is orthogonal to ll and is pointing outward PiP_{i} at xx.

Meanwhile, since yiPi̊y_{i}\in\mathring{P_{i}}, we definitely have

yi,N¯i=0,\displaystyle\langle-y_{i},\bar{N}_{i}\rangle=0,
yi,N¯j>0for ji.\displaystyle\langle-y_{i},\bar{N}_{j}\rangle>0\quad\text{for }j\neq i.

Recall that xP1P2x\in P_{1}\cap P_{2} and hence x,N¯i=0\langle x,\bar{N}_{i}\rangle=0 for each ii, we thus obtain

xyi,N¯i=0,\displaystyle\langle x-y_{i},\bar{N}_{i}\rangle=0,
xyi,N¯j>0,for ji.\displaystyle\langle x-y_{i},\bar{N}_{j}\rangle>0,\quad\text{for }j\neq i.

Combining all above and invoking that ν(x)=c1N¯1c2N¯2\nu(x)=-c_{1}\bar{N}_{1}-c_{2}\bar{N}_{2}, we thus find: c1<0,c2<0c_{1}<0,c_{2}<0. This proves the Claim and completes the proof.

Proof of (38)\Leftrightarrow(4): Using

cosθ2=cos(θ2θ1)cosθ1sin(θ2θ1)sinθ1,\displaystyle\cos\theta^{2}=\cos(\theta^{2}-\theta^{1})\cos\theta^{1}-\sin(\theta^{2}-\theta^{1})\sin\theta^{1},

we see that (38) is equivalent to

(cosη2+cosη1cosα)sinθ1\displaystyle(\cos\eta^{2}+\cos\eta^{1}\cos\alpha)\sin\theta^{1}
\displaystyle\geq (cos(θ2θ1)cosθ1sin(θ2θ1)sinθ1+cosθ1cosα)sinη1.\displaystyle\Big{(}\cos(\theta^{2}-\theta^{1})\cos\theta^{1}-\sin(\theta^{2}-\theta^{1})\sin\theta^{1}+\cos\theta^{1}\cos\alpha\Big{)}\sin\eta^{1}.

After rearranging, we get

sinθ1cosη2+cosαsin(θ1η1)\displaystyle\sin\theta^{1}\cos\eta^{2}+\cos\alpha\sin{(\theta^{1}-\eta^{1})}
=\displaystyle= sinθ1cosη2+cosα(sinθ1cosη1sinη1cosθ1)\displaystyle\sin\theta^{1}\cos\eta^{2}+\cos\alpha(\sin\theta^{1}\cos\eta^{1}-\sin\eta^{1}\cos\theta^{1})
\displaystyle\geq cos(θ2θ1)(cosθ1sinη1sinθ1cosη1)\displaystyle\cos(\theta^{2}-\theta^{1})(\cos\theta^{1}\sin\eta^{1}-\sin\theta^{1}\cos\eta^{1})
+sinθ1(cos(θ2θ1)cosη1sin(θ2θ1)sinη1)\displaystyle+\sin\theta^{1}\Big{(}\cos{(\theta^{2}-\theta^{1})}\cos\eta^{1}-\sin(\theta^{2}-\theta^{1})\sin\eta^{1}\Big{)}
=\displaystyle= cos(θ2θ1)sin(θ1η1)+sinθ1cos(θ2θ1+η1).\displaystyle-\cos(\theta^{2}-\theta^{1})\sin(\theta^{1}-\eta^{1})+\sin\theta^{1}\cos(\theta^{2}-\theta^{1}+\eta^{1}).

which is just (4).∎

References