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

Some rigidity results on compact hypersurfaces with capillary boundary in the hyperbolic space

Yimin Chen Department of Mathematics
Pusan National University
Busan 46241
Republic of Korea
sherlockpoe@pusan.ac.kr
 and  Juncheol Pyo Department of Mathematics
Pusan National University
Busan 46241
Republic of Korea
jcpyo@pusan.ac.kr
Abstract.

In this paper, we prove a Heintze-Karcher type inequality for capillary hypersurfaces supported on various hypersurfaces in the hyperbolic space. The equality case only occurs on capillary totally umbilical hypersurfaces. Then we apply this result to prove the Alexandrov type theorem for embedded capillary hypersurfaces in the hyperbolic space. In addition, we prove some other rigidity results for capillary hypersurfaces supported on totally geodesic plane in 𝔹+n+1\mathbb{B}^{n+1}_{+}.

1. Introduction

Let n+1\mathbb{H}^{n+1} be the (n+1)(n+1)-dimensional hyperbolic space. We will use both the Poincaré ball model and the Poincaré half space model of n+1\mathbb{H}^{n+1} throughout this paper. Let δ\delta be the Euclidean metric and ||=δ(,)1/2|\cdot|=\delta(\cdot,\cdot)^{1/2} be the Euclidean norm. The Poincaré ball model is defined as follows:

(𝔹n+1,4(1|x|2)2δ),\displaystyle\left(\mathbb{B}^{n+1},\frac{4}{(1-|x|^{2})^{2}}\delta\right),

where 𝔹n+1\mathbb{B}^{n+1} is the unit Euclidean ball centered at the origin.

The Poincaré half space model is defined as follows:

(+n+1,1xn+12δ),\displaystyle\left(\mathbb{R}^{n+1}_{+},\frac{1}{x_{n+1}^{2}}\delta\right),

where +n+1\mathbb{R}^{n+1}_{+} is the half Euclidean space on which the last coordinate function xn+1x_{n+1} is strictly positive.

Let SS be an umbilical hypersurface in n+1\mathbb{H}^{n+1} with its principal curvature λ\lambda, and Σn+1\Sigma\hookrightarrow\mathbb{H}^{n+1} is an immersed nn-dimensional manifold into n+1\mathbb{H}^{n+1} such that x(Σ)Sx(\partial\Sigma)\subset S. If Σ\Sigma intersects SS at a constant angle θ\theta, Σ\Sigma is said to be a capillary hypersurface. In particular if θ=π2\theta=\frac{\pi}{2}, Σ\Sigma is said to be a free boundary hypersurface. Furthermore, we call SS the supporting hypersurface of the capillary hypersurface Σ\Sigma. The study of capillary hypersurfaces has a long history since the works by Young in [27] and Laplace in [12]. The notable result of Fraser and Schoen in [4] reveals the relation between the Steklov eigenvalue and the free boundary minimal hypersurfaces in the unit ball. Based on their significant work, properties of free boundary hypersurfaces have been established, including some geometric inequalities.

There are several results on geometric inequalities on capillary hypersurfaces in a space form. In [22] and [26], Scheuer, Wang, Weng and Xia have established a family of Alexandrov-Fenchel’s type inequalities on capillary hypersurfaces in a geodesic ball. In [24] and [7], the authors give a complete family of Alexandrov-Fenchel’s type inequalities on capillary hypersurfaces in Euclidean half space. And in [3], Chen, Hu and Li proved Perez type inequality on free boundary hypersurfaces in a geodesic ball.

In [21], Ros used Reilly’s formula (see [20]) to prove the following result.

Theorem A.

Let Ω\Omega be an (n+1)(n+1)-dimensional Riemannian manifold with boundary and non-negative Ricci curvature. Suppose the boundary M=ΩM=\partial\Omega is mean convex (i.e. the mean curvature HM>0H_{M}>0), then

M1HM𝑑A(n+1)V(Ω),\int_{M}\frac{1}{H_{M}}dA\geq(n+1)V(\Omega),

where V(Ω)V(\Omega) denotes the volume of Ω\Omega. Moreover, the equality holds if and only if Ω\Omega is isometric to a Euclidean ball.

There are several results on (weighted) Heintze-Karcher type inequalities for embedded closed hypersurfaces in different Riemannian manifolds. In [2], Brendle proved a weighted Heintze-Karcher type inequality for embedded, closed mean convex hypersurfaces in a family of warped product spaces. In [19], Qiu and Xia proved a weighted Heintze-Karcher type inequality for the case in which the ambient space is a Riemannian manifold with a negative lower bound on its sectional curvature. In [13], Li and Xia proved a weighted Heintze-Karcher type inequality when the ambient space is a sub-static manifold.

For free boundary hypersurfaces, the second author proved a Heintze-Karcher type inequality for those supported on totally geodesics hyperplane in a space form in [18]. Guo and Xia proved the inequality for those supported on horospheres and equidistant hypersurfaces in [6]. In [25], Wang and Xia prove the case for those supported on geodesic balls.

Recently in the paper [10], Jia, Xia and Zhang proved a Heintze-Karcher type inequality for capillary hypersurfaces in Euclidean half space. Furthermore, they extended their results in Euclidean wedge and anisotropic half space (See [8] and [9]). Inspired by their result, we will prove a Heintze-Karcher type inequality for capillary hypersurfaces in n+1\mathbb{H}^{n+1}.

In Poincaré ball model (𝔹n+1,4(1|x|2)2δ)(\mathbb{B}^{n+1},\frac{4}{(1-|x|^{2})^{2}}\delta), let Σ\Sigma be a compact, embedded capillary hypersurface in

𝔹+n+1={x𝔹n+1:δ(x,En+1)0}\mathbb{B}^{n+1}_{+}=\{x\in\mathbb{B}^{n+1}:\delta(x,E_{n+1})\geq 0\}

supported on a totally geodesic plane

P={x𝔹n+1:δ(x,En+1)=0}.P=\{x\in\mathbb{B}^{n+1}:\delta(x,E_{n+1})=0\}.

Denote Ω\Omega the region enclosed by Σ\Sigma and PP, and TPT\subset P the region in PP with Ω=ΣT\partial\Omega=\Sigma\cup T. Denote V0=1+|x|21|x|2V_{0}=\frac{1+|x|^{2}}{1-|x|^{2}}, we have the following Heintze-Karcher type inequality.

Theorem 1.

Let Σ𝔹+n+1\Sigma\subset\mathbb{B}^{n+1}_{+} be a compact, embedded capillary hypersurface supported on PP. Let Ω\Omega be a domain enclosed by Σ\Sigma and TPT\subset P. If the contact angle θ(0,π2]\theta\in(0,\frac{\pi}{2}], and the mean curvature H1>0H_{1}>0 on Σ\Sigma, we have

(1) ΣV0H1𝑑A(n+1)ΩV0𝑑Ω+ncotθ(TV0𝑑AT)2ΣV0𝑑s.\displaystyle\int_{\Sigma}\frac{V_{0}}{H_{1}}dA\geq(n+1)\int_{\Omega}V_{0}d\Omega+n\cot\theta\frac{\left(\int_{T}V_{0}dA_{T}\right)^{2}}{\int_{\partial\Sigma}V_{0}ds}.

In addition, the equality holds if and only if Σ\Sigma is umbilical.

Next, we consider Poincaré half space model (+n+1,1xn+12δ)(\mathbb{R}^{n+1}_{+},\frac{1}{x_{n+1}^{2}}\delta) of the hyperbolic space. Denote the position vector in Poincaré half space model by x~\widetilde{x}. Fixing aa a vector field in n+1\mathbb{H}^{n+1} which is constant with respect to (n+1,δ+)(\mathbb{R}^{n+1},\delta_{+}) and 0<ϕ<π20<\phi<\frac{\pi}{2}, we consider a umbilical hypersurface Lϕ,aL_{\phi,a} defined by

(2) Lϕ,a={x~+n+1:xn+11=tanϕxa},L_{\phi,a}=\{\widetilde{x}\in\mathbb{R}^{n+1}_{+}:x_{n+1}-1=\tan\phi x_{a}\},

where xa=δ(x~,a)x_{a}=\delta(\widetilde{x},a) and aa is a vector field in n+1\mathbb{H}^{n+1} which is constant with respect to (+n+1,δ)(\mathbb{R}^{n+1}_{+},\delta) and satisfies δ(a,En+1)=0\delta(a,E_{n+1})=0 and δ(a,a)=1\delta(a,a)=1.It is well known that Lϕ,aL_{\phi,a} is a umbilical hypersurface with principal curvature 0<cosϕ10<\cos\phi\leq 1. We define Bϕ,aintB^{\mathrm{int}}_{\phi,a} as follows:

Bϕ,aint={x~n+1:xn+11tanϕxa}.B^{\mathrm{int}}_{\phi,a}=\{\widetilde{x}\in\mathbb{H}^{n+1}:x_{n+1}-1\geq\tan\phi x_{a}\;\}.

In BϕintB_{\phi}^{\mathrm{int}}, we consider compact embedded capillary hypersurfaces supported on Lϕ,aL_{\phi,a}.

Analogously, denote Vn+1=1xn+1V_{n+1}=\frac{1}{x_{n+1}} and we have the following Heintze-Karcher type inequality

Theorem 2.

Let 0ϕ<π20\leq\phi<\frac{\pi}{2} and ΣBϕ,aint\Sigma\subset B^{\mathrm{int}}_{\phi,a} be a compact, embedded capillary hypersurface. The supporting hypersurface is LϕL_{\phi}. Let Ω\Omega be the domain bounded by Σ\Sigma and TLϕ,aT\subset L_{\phi,a}. If the contact angle θ(0,π2]\theta\in(0,\frac{\pi}{2}], and the mean curvature H1>0H_{1}>0 on Σ\Sigma, we have

(3) ΣVn+1H1𝑑A(n+1)ΩVn+1𝑑Ω+ncosθ(TVn+1𝑑AT)2Σg¯(cosϕx~+sinϕa,μ)𝑑s\displaystyle\int_{\Sigma}\frac{V_{n+1}}{H_{1}}dA\geq(n+1)\int_{\Omega}V_{n+1}d\Omega+n\cos\theta\frac{\left(\int_{T}V_{n+1}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\phi a,\mu)ds}

where μ\mu is the outward unit conormal vector field of Σ\partial\Sigma in Σ\Sigma. In addition, the equality holds if and only if Σ\Sigma is umbilical.

The classical Alexandrov type theorem says that the geodesic balls are the only closed embedded hypersurfaces with constant mean curvature (CMC) in space form. It has been proved using a method called moving planes (see [1]).

Another powerful tool to prove the Alexandrov type theorem is the method provided by Ros. In 1986, Ros used a Heintze-Karcher type formula in [21] to prove a higher-order version of an Alexandrov type theorem, which says that embedded hypersurfaces with constant kk-th mean curvature in Euclidean space can only be a sphere. For the general ambient space, Brendle [2] proved that any closed embedded CMC hypersurface in a family of space warped product spaces can only be a slice.

As we mentioned above, there exists certain Heintze-Karcher type inequalities on free boundary hypersurfaces on different supporting hypersurfaces, Wang and Xia [25] proved the Alexandrov type theorem for free boundary hypersurface supported on geodesic spheres in space forms. Moreover, for free boundary hypersurfaces supported on horospheres and equidistant hypersurfaces, we refer to Guo and Xia [6]. The second author [18] also studied the case of free boundary hypersurfaces supported on totally geodesic hyperplane in space forms. For capillary hypersurface, Jia, Xia and Zhang (see [10]) prove an Alexandrov type theorem for capillary hypersurfaces in the Euclidean half space and Euclidean ball. Jia, Wang, Xia and Zhang prove the capillary case of the support hypersurface being Euclidean wedges and the geodesic plane in anisotropic Euclidean half space (see [8] and [9]). In this paper, we will prove the following Alexandrov type theorems for capillary hypersurfaces in the hyperbolic space.

Theorem 3.

Let Σ\Sigma be a compact embedded CMC hypersurface in 𝔹+n+1\mathbb{B}^{n+1}_{+} supported on a the totally geodesic plane PP, and the contact angle satisfies θ(0,π2]\theta\in(0,\frac{\pi}{2}]. Then Σ\Sigma is totally umbilical except for being totally geodesic.

Theorem 4.

Let Σ\Sigma be a compact embedded CMC capillary hypersurface in Bϕ,aintB^{\mathrm{int}}_{\phi,a} supported on Lϕ,aL_{\phi,a}, and the contact angle satisfies θ(0,π2]\theta\in(0,\frac{\pi}{2}]. In addition, we assume ϕ+θ>π2\phi+\theta>\frac{\pi}{2} for ϕ>0\phi>0 or θ=π2\theta=\frac{\pi}{2} for ϕ=0\phi=0. Then Σ\Sigma is umbilical except for being totally geodesic.

Remark.

The assumption “ ϕ+θ>π2\phi+\theta>\frac{\pi}{2} for ϕ>0\phi>0 or θ=π2\theta=\frac{\pi}{2} for ϕ=0\phi=0” is only used for guaranteeing the existence of the convex point, which can be substituted by assuming that the constant mean curvature H>0H>0, see Corollary 1.

Intriguingly, we also prove some other rigidity results on immersed capillary hypersurfaces supported on a totally geodesic plane PP in the hyperbolic space. Koh and Lee [11] used the Minkowski identity for closed hypersurfaces in Euclidean space to classify the immersed hypersurfaces with constant ratio of higher-order mean curvatures in Euclidean space. Since we will give a Minkowski type formula on capillary hypersurfaces in the Section 3, we can prove the following result:

Theorem 5.

Let Σ𝔹+n+1\Sigma\subset\mathbb{B}^{n+1}_{+} be an immersed capillary hypersurface supported on PP. Suppose k>l1k>l\geq 1. If there exists a constant α\alpha such that on Σ\Sigma,

HkHl=α,Hl>0,\displaystyle\frac{H_{k}}{H_{l}}=\alpha,\quad H_{l}>0,

we have that Σ\Sigma is totally umbilical except for being totally geodesic.

Remark.

We note that the second author proved the results in Theorem 1, Theorem 3 and Theorem 5 for the free boundary case in [18]. Guo and Xia proved the result in Theorem 2 and Theorem 4 for the free boundary case in [6].

Let On+1O\in\mathbb{H}^{n+1} be a point and r(p)=dist(p,O)r(p)=\mbox{dist}(p,O) be the distance function in the hyperbolic space. The metric in the hyperbolic space can also be written as the following warped product form

g¯=dr2+sinhrgSn.\overline{g}=dr^{2}+\sinh rg_{S^{n}}.

It can be easily observed that the position vector field xx in Pioncaré ball model satisfies x=sinhrrx=\sinh r\partial r if we choose OO be the origin of 𝔹n+1\mathbb{B}^{n+1}. A hypersurface MM in n+1\mathbb{H}^{n+1} is called star-shaped if g¯(x,ν)=g¯(sinhrr,ν)>0\overline{g}(x,\nu)=\overline{g}(\sinh r\partial r,\nu)>0. It has been proved that in some specific warped product manifolds (including the hyperbolic space), an immersed, compact, closed, star-shaped hypersurface with constant mean curvature must be a geodesic sphere (See [16, Corollary 5]). Now we give a version of capillary hypersurfaces in the hyperbolic space.

Theorem 6.

Let Σ𝔹+n+1\Sigma\hookrightarrow\mathbb{B}^{n+1}_{+} be a compact, immersed CMC capillary hypersurface supported on PP. If Σ\Sigma is star-shaped, Σ\Sigma is totally umbilical.

Since xx is embedding immediately if it is star-shaped, Theorem 3 implies Theorem 6 when θ(0,π2]\theta\in(0,\frac{\pi}{2}].

The remainder of this paper is organized as follows. In Section 2, we collect basic facts on the hyperbolic space. In Section 3, we prove Minkowski type formulae. In Section 4, we give proofs of Theorem 1 and Theorem 3. In Section 5, we give proofs of Theorem 2 and Theorem 4. In Section 6, we consider case of supporting hypersurfaces being geodesic spheres. Finally, in Section 7, we prove two other rigidity results for the capillary hypersurfaces supported on a totally geodesic plane in the hyperbolic space.

Acknowledgements.

This work is supported by the National Research Foundation of Korea (grant NRF No.2021R1A4A1032418). We would also like to thank Prof. Haizhong Li from Tsinghua University, Prof. Yingxiang Hu from Beihang University and Prof. Chao Xia from Xiamen University for their useful comments and constant supports.

2. Preliminaries

2.1. Capillary hypersurfaces and properties of the higher-order mean curvature HrH_{r}

Let g¯\overline{g} and ¯\overline{\nabla} be the hyperbolic metric and the Levi-Civita connection of n+1\mathbb{H}^{n+1}, respectively. Let X:Σ¯BX:\Sigma\rightarrow\mathbb{\overline{}}B be an immersion of an nn-dimensional Riemannian manifold Σ\Sigma with boundary Σ\partial\Sigma into an (n+1)(n+1)-dimensional Riemannian manifold BB with boundary B\partial B. Σ\Sigma is called a capillary hypersurface in BB if the immersion satisfies the following

X(int(Σ))int(B),X(Σ)B,\displaystyle X(\mbox{int}(\Sigma))\subset\mbox{int}(B),\quad X(\partial\Sigma)\subset\partial B,

and Σ\Sigma and B\partial B intersects with a constant contact angle θ(0,π)\theta\in(0,\pi) along Σ\partial\Sigma. On Σ\partial\Sigma, there exists four unit normal vector fields, {ν,μ,N¯,ν¯}\{\nu,\,\mu,\,\overline{N},\,\overline{\nu}\}, on a 2-dimensional subspace of Tpn+1T_{p}\mathbb{H}^{n+1} as follows:

  1. (i)

    ν\nu is the outward unit normal vector field of isometric immersion X:ΣBX:\Sigma\rightarrow B.

  2. (ii)

    μ\mu is the outward unit normal vector field of isometric immersion y:ΣΣy:\partial\Sigma\rightarrow\Sigma.

  3. (iii)

    N¯\overline{N} is the outward unit normal vector field of isometric immersion z:BBz:\partial B\rightarrow B.

  4. (iv)

    ν¯\overline{\nu} is the outward unit normal vector field of isometric immersion w:ΣBw:\partial\Sigma\rightarrow\partial B.

The second fundamental form of x:Σn+1x:\Sigma\rightarrow\mathbb{H}^{n+1} denoted by hh is defined as follows:

h(Y,Z)=g¯(¯Yν,Z), for Y,ZT(Σ),\displaystyle h(Y,Z)=\overline{g}(\overline{\nabla}_{Y}\nu,Z),\mbox{ for }Y,\;Z\in T(\Sigma),

and the second fundamental form of w:ΣBw:\partial\Sigma\rightarrow\partial B denoted by hΣh^{\partial\Sigma} is defined as follows:

hΣ(Y,Z)=g¯(¯Yν¯,Z), for Y,ZT(Σ).\displaystyle h^{\partial\Sigma}(Y,Z)=\overline{g}(\overline{\nabla}_{Y}\overline{\nu},Z),\mbox{ for }Y,\;Z\in T(\partial\Sigma).
ν\displaystyle\nu      μ\muN¯\displaystyle\overline{N}ν¯\displaystyle\overline{\nu}Σ\displaystyle\SigmaB\displaystyle\partial BB\displaystyle Bθ\displaystyle\theta
Figure 1. Capillary hypersurface supported on B\partial B

Now we have the following relation of these vector fields,

(6) {μ=sinθN¯+cosθν¯,ν=cosθN¯+sinθν¯,\displaystyle\left\{\begin{array}[]{l}\mu=\sin\theta\overline{N}+\cos\theta\overline{\nu},\\ \nu=-\cos\theta\overline{N}+\sin\theta\overline{\nu},\end{array}\right.

where θ\theta denotes the angle between N¯\overline{N} and ν-\nu. The following lemma is well-known and fundamental.

Lemma 1.

Let X:ΣBn+1X:\Sigma\rightarrow B\subset\mathbb{H}^{n+1} be an isometric immersion of a capillary hypersurface supported on the totally umbilical hypersurface B\partial B. Then μ\mu is a principal direction of Σ\Sigma, that is,

(7) ¯μν=h(μ,μ)μ.\displaystyle\overline{\nabla}_{\mu}\nu=h(\mu,\mu)\mu.
Proof.

The proof can be found in [25, Proposition 2.1], we give a proof for completeness. Without loss of generality, let {eα}\{e_{\alpha}\} be the orthonormal basis along Σ\partial\Sigma and α=1,2,,n1\alpha=1,2,\dots,n-1. Then, we have the following

¯μν=\displaystyle\overline{\nabla}_{\mu}\nu= g¯(¯μν,μ)μ+α=1n1g¯(¯μν,eα)eα\displaystyle\overline{g}(\overline{\nabla}_{\mu}\nu,\mu)\mu+\sum_{\alpha=1}^{n-1}\overline{g}(\overline{\nabla}_{\mu}\nu,e_{\alpha})e_{\alpha}
=\displaystyle= h(μ,μ)μα=1n1g¯(¯eα(cosθN¯+sinθν¯),sinθN¯+cosθν¯)eα\displaystyle h(\mu,\mu)\mu-\sum_{\alpha=1}^{n-1}\overline{g}(\overline{\nabla}_{e_{\alpha}}(-\cos\theta\overline{N}+\sin\theta\overline{\nu}),\sin\theta\overline{N}+\cos\theta\overline{\nu})e_{\alpha}
=\displaystyle= h(μ,μ)μhΣ(eα,ν¯)eα\displaystyle h(\mu,\mu)\mu-h^{\partial\Sigma}(e_{\alpha},\overline{\nu})e_{\alpha}
=\displaystyle= h(μ,μ)μ,\displaystyle h(\mu,\mu)\mu,

where the last equality holds because B\partial B is totally umbilical. ∎

Let hij=h(ei,ej)h_{ij}=h(e_{i},e_{j}) and λ=(λ1,λ2,,λn)\lambda=(\lambda_{1},\lambda_{2},\dots,\lambda_{n}) be the eigenvalues of the matrix (hij)i,j=1n(h_{ij})_{i,j=1}^{n}, that is, the principal curvatures of Σ\Sigma. Let k{1,2,,n}k\in\{1,2,\dots,n\}, the kk-th mean curvature is defined by

Sk=1k!1i1<<iknλ1λ2λik,\displaystyle S_{k}=\frac{1}{k!}\sum\limits_{1\leq i_{1}<\cdots<i_{k}\leq n}\lambda_{1}\lambda_{2}\dots\lambda_{i_{k}},

and the normalized kk-th mean curvature is defined by Hk=(nk)1SkH_{k}=\binom{n}{k}^{-1}S_{k}. The associated Newton transformation is defined by induction,

T0=I,\displaystyle T_{0}=I,
Tk=SkITk1h.\displaystyle T_{k}=S_{k}I-T_{k-1}\circ h.

Using the induction formulae, we have the following properties.

tr(Tk)=(nk)Sk=(nk)(nk)Hk,\displaystyle\mbox{tr}(T_{k})=(n-k)S_{k}=(n-k)\binom{n}{k}H_{k},
tr(Tkh)=i,j=1n(Tk)jihij=(k+1)Sk+1=(nk)(nk)Hk+1.\displaystyle\mbox{tr}(T_{k}\circ h)=\sum_{i,j=1}^{n}(T_{k})_{j}^{i}h_{i}^{j}=(k+1)S_{k+1}=(n-k)\binom{n}{k}H_{k+1}.

Let Γk+={λn:Hi>0,i=1,2,,k}\Gamma^{+}_{k}=\{\lambda\in\mathbb{R}^{n}:H_{i}>0,i=1,2,\dots,k\}. Then, for 1l<kn1\leq l<k\leq n and λΓl+\lambda\in\Gamma^{+}_{l}, we have the following classic Newton-Maclaurin inequality,

(8) HlHl1HkHk1.\displaystyle\frac{H_{l}}{H_{l-1}}\geq\frac{H_{k}}{H_{k-1}}.

The equality in (8) holds if and only if λ=c\lambda=c\mathcal{I}, where c>0c>0 is a constant and =(1,1,,1)n\mathcal{I}=(1,1,\dots,1)\in\mathbb{R}^{n}.

2.2. Properties in the Poincaré ball model

Let xx be the position vector in 𝔹n+1\mathbb{B}^{n+1}. The following fact is well-known:

(9) ¯x=V0g¯,\displaystyle\overline{\nabla}x=V_{0}\overline{g},

where V0=1+|x|21|x|2V_{0}=\frac{1+|x|^{2}}{1-|x|^{2}} is a smooth function in n+1\mathbb{H}^{n+1} satisfying

¯2V0=V0g¯.\overline{\nabla}^{2}V_{0}=V_{0}\overline{g}.

A vector field XX is called a conformal Killing vector field in (M¯,g¯)(\overline{M},\overline{g}), if the Lie derivative \mathcal{L} satisfies

Xg¯=fg¯.\displaystyle\mathcal{L}_{X}\overline{g}=f\overline{g}.

In particular, XX is call a Killing vector if f=0f=0. In this case, the key Killing vector field YaY_{a} in Poincaré ball model we use is

Ya=12(1+|x|2)aδ(x,a)x,\displaystyle Y_{a}=\frac{1}{2}(1+|x|^{2})a-\delta(x,a)x,

where aa is an arbitrary vector field in n+1\mathbb{H}^{n+1} which is constant with respect to (𝔹n+1,δ)(\mathbb{B}^{n+1},\delta). It is given by Wang and Xia in [25].

Let E1,E2,,En+1E_{1},E_{2},\dots,E_{n+1} form the coordinate frame in the conformal Euclidean metric, and the corresponding normalized vectors E¯1,E¯2,,E¯n+1\overline{E}_{1},\overline{E}_{2},\dots,\overline{E}_{n+1} form an orthonormal frame on (𝔹n+1,g¯)(\mathbb{B}^{n+1},\overline{g}). For A,B{1,2,,n+1}A,B\in\{1,2,\dots,n+1\}, it holds that

(10) 12(g¯(¯E¯AYa,E¯B)+g¯(¯E¯BYa,E¯A))=0.\displaystyle\frac{1}{2}\left(\overline{g}(\overline{\nabla}_{\overline{E}_{A}}Y_{a},\overline{E}_{B})+\overline{g}(\overline{\nabla}_{\overline{E}_{B}}Y_{a},\overline{E}_{A})\right)=0.

This can be verified by the following properties.

Proposition 1 (See [25]).

Let a=i=1n+1aiEia=\sum\limits_{i=1}^{n+1}a_{i}E_{i} be an arbitrary vector field satisfying that aia_{i} are constant and δ(a,En+1)=0\delta(a,E_{n+1})=0. Denote the normalized vector of aa and En+1E_{n+1} in (𝔹+n+1,4(1|x|2)2δ)(\mathbb{B}^{n+1}_{+},\frac{4}{(1-|x|^{2})^{2}}\delta) by a¯\bar{a} and E¯n+1\bar{E}_{n+1} respectively, that is, a¯=1|x|22a\bar{a}=\frac{1-|x|^{2}}{2}a and E¯n+1=1|x|22En+1\bar{E}_{n+1}=\frac{1-|x|^{2}}{2}E_{n+1}. Then, for any ZTn+1Z\in T\mathbb{H}^{n+1}, we have the following:

  1. (a)

    ¯Za=g¯(Z,x)a¯+g¯(a¯,x)Zg¯(Z,a¯)x\overline{\nabla}_{Z}a=\overline{g}(Z,x)\bar{a}+\overline{g}(\bar{a},x)Z-\overline{g}(Z,\bar{a})x,

  2. (b)

    ¯Z(1|x|22a)=1|x|22[g¯(x,a¯)Zg¯(Z,a¯)x]\overline{\nabla}_{Z}(\frac{1-|x|^{2}}{2}a)=\frac{1-|x|^{2}}{2}\left[\overline{g}(x,\bar{a})Z-\overline{g}(Z,\bar{a})x\right],

  3. (c)

    ¯ZV0=g¯(x,Z)\overline{\nabla}_{Z}V_{0}=\overline{g}(x,Z),

  4. (d)

    ¯ZYa=g¯(x,Z)a¯g¯(Z,a¯)x\overline{\nabla}_{Z}Y_{a}=\overline{g}(x,Z)\bar{a}-\overline{g}(Z,\bar{a})x.

2.3. Properties in the Poincaré half space model

In Poincaré half space model n+1=(+n+1,1xn+12δ)\mathbb{H}^{n+1}=(\mathbb{R}^{n+1}_{+},\frac{1}{x_{n+1}^{2}}\delta), we denote x~\widetilde{x} the position vector field in +n+1\mathbb{R}^{n+1}_{+}.

Let E1,E2,,En+1E_{1},E_{2},\dots,E_{n+1} form the coordinate frame in the Euclidean metric, and the normalized vectors E¯1,E¯2,,E¯n+1\overline{E}_{1},\overline{E}_{2},\dots,\overline{E}_{n+1} form an orthonormal frame on (+n+1,g¯)(\mathbb{R}^{n+1}_{+},\overline{g}). A conformal Killing vector field given by

Xn+1:=x~En+1X_{n+1}:=\widetilde{x}-E_{n+1}

plays an important role in [6] and [5]. For A,B{1,2,,n+1}A,B\in\{1,2,\dots,n+1\}, it holds that

(11) 12(g¯(¯E¯AXn+1,E¯B)+g¯(¯E¯BXn+1,E¯A))=Vn+1g¯,\displaystyle\frac{1}{2}\left(\overline{g}(\overline{\nabla}_{\overline{E}_{A}}X_{n+1},\overline{E}_{B})+\overline{g}(\overline{\nabla}_{\overline{E}_{B}}X_{n+1},\overline{E}_{A})\right)=V_{n+1}\overline{g},

where Vn+1=1xn+1V_{n+1}=\frac{1}{x_{n+1}} is a smooth function satisfying

¯2Vn+1=Vn+1g¯.\overline{\nabla}^{2}V_{n+1}=V_{n+1}\overline{g}.

Similarly, in the Poincaré half space model, we have the following proposition, the proof of which is omitted.

Proposition 2 (See [5]).

Let a=i=1n+1aiEia=\sum\limits_{i=1}^{n+1}a_{i}E_{i} be an arbitrary vector field satisfying that aia_{i} are constant and δ(a,En+1)=0\delta(a,E_{n+1})=0. Denote the normalized vector of aa and En+1E_{n+1} in (+n+1,1xn+12δ)(\mathbb{R}^{n+1}_{+},\frac{1}{x_{n+1}^{2}}\delta) by a¯\bar{a} and E¯n+1\bar{E}_{n+1} respectively, that is, a¯=xn+1a\bar{a}=x_{n+1}a and E¯n+1=xn+1En+1\bar{E}_{n+1}=x_{n+1}E_{n+1}. Then, for any YTn+1Y\in T\mathbb{H}^{n+1}, we have the following:

  1. (a)

    ¯Yx~=g¯(Y,E¯n+1)x~+g¯(Y,x~)E¯n+1\overline{\nabla}_{Y}\widetilde{x}=-\overline{g}(Y,\overline{E}_{n+1})\widetilde{x}+\overline{g}(Y,\widetilde{x})\overline{E}_{n+1},

  2. (b)

    ¯Ya=g¯(Y,E¯n+1)a¯+g¯(Y,a¯)En+1\overline{\nabla}_{Y}a=-\overline{g}(Y,\overline{E}_{n+1})\overline{a}+\overline{g}(Y,\overline{a})E_{n+1},

  3. (c)

    ¯YEn+1=1xn+1Y\overline{\nabla}_{Y}E_{n+1}=-\frac{1}{x_{n+1}}Y,

  4. (d)

    ¯YE¯n+1=g¯(E¯n+1,Y)E¯n+1Y\overline{\nabla}_{Y}\overline{E}_{n+1}=\overline{g}(\overline{E}_{n+1},Y)\overline{E}_{n+1}-Y.

It is easy to see that (11) can be obtained from (a) and (c) in Proposition 2. Moreover, from (a) and (b) in Proposition 2, we have the following:

(12) 12(g¯(¯E¯Aa,E¯B)+g¯(¯E¯Ba,E¯A))=0,\displaystyle\frac{1}{2}\left(\overline{g}(\overline{\nabla}_{\overline{E}_{A}}a,\overline{E}_{B})+\overline{g}(\overline{\nabla}_{\overline{E}_{B}}a,\overline{E}_{A})\right)=0,

and

(13) 12(g¯(¯E¯Ax~,E¯B)+g¯(¯E¯Bx~,E¯A))=0,\displaystyle\frac{1}{2}\left(\overline{g}(\overline{\nabla}_{\overline{E}_{A}}\widetilde{x},\overline{E}_{B})+\overline{g}(\overline{\nabla}_{\overline{E}_{B}}\widetilde{x},\overline{E}_{A})\right)=0,

that is, aa and x~\widetilde{x} are Killing vector fields in (+n+1,1xn+12δ)(\mathbb{R}^{n+1}_{+},\frac{1}{x_{n+1}^{2}}\delta).

3. Minkowski type formulae in the hyperbolic space

In [25], the authors give a Minkowski type formula for capillary hypersurface MM in a geodesic ball of radius RR in a space form 𝕄n+1(K)\mathbb{M}^{n+1}(K) (K=0,1, 1K=0,\,-1,\,1),

M(n(Va+snK(R)cosθg¯(Ya,ν))S1g¯(Xa,ν))𝑑A=0,\displaystyle\int_{M}\left(n(V_{a}+\mbox{sn}_{K}(R)\cos\theta\overline{g}(Y_{a},\nu))-S_{1}\overline{g}(X_{a},\nu)\right)dA=0,

where

snK(r)={r, when K=0;sinr, when K=1;sinhr, when K=1,\displaystyle\mbox{sn}_{K}(r)=\left\{\begin{array}[]{ll}r,&\mbox{ when }K=0;\\ \sin r,&\mbox{ when }K=1;\\ \sinh r,&\mbox{ when }K=-1,\\ \end{array}\right.

and XaX_{a} is a conformal Killing vector field tangent to the geodesic sphere of radius RR, which is the supporting hypersurface of MM. We introduce our results on the case (supported on geodesic sphere in n+1\mathbb{H}^{n+1}) in the Section 6.

3.1. A Minkowski type formula for capillary hypersurfaces in Poincaré ball model

In 𝔹+n+1\mathbb{B}^{n+1}_{+}, the outer unit normal vector field along PP is N¯=1|x|22En+1\overline{N}=-\frac{1-|x|^{2}}{2}E_{n+1}. We note that the position vector field xx in the Poincaré ball model is tangent to PP, therefore g¯(x,N¯)=0\overline{g}(x,\overline{N})=0.

ν\displaystyle\nuν¯\displaystyle\overline{\nu}N¯\displaystyle\overline{N}μ\displaystyle\muEn+1\displaystyle E_{n+1}Σ\displaystyle\SigmaP\displaystyle PO\displaystyle O
Figure 2. Capillary hypersurface Σ\Sigma supported on a totally geodesic plane PP.

Let Yn+1=YEn+1=12(1+|x|2)En+1δ(x,En+1)xY_{n+1}=Y_{E_{n+1}}=\frac{1}{2}(1+|x|^{2})E_{n+1}-\delta(x,E_{n+1})x and En+1E_{n+1} be the unit constant vector with respect to δ\delta. Then, we can derive a Minkowski type formula as follows.

Proposition 3.

Let x:Σ𝔹+n+1x:\Sigma\subset\mathbb{B}^{n+1}_{+} be a compact, immersed capillary hypersurface supported on PP, intersecting PP at a constant angle θ\theta. For k{1,2,,n}k\in\{1,2,\dots,n\}, we have

(14) Σ[Hk1(V0cosθg¯(Yn+1,ν))Hkg¯(x,ν)]𝑑A=0.\displaystyle\int_{\Sigma}\left[H_{k-1}(V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu))-H_{k}\overline{g}(x,\nu)\right]dA=0.
Proof.

Restricting (9) to Σ\Sigma and let {ei}i=1n\{e_{i}\}_{i=1}^{n} be the orthonormal frame on Σ\Sigma, we have

(15) g¯(eixT,ej)=V0δijhijg¯(x,ν),\displaystyle\overline{g}(\nabla_{e_{i}}x^{T},e_{j})=V_{0}\delta_{ij}-h_{ij}\overline{g}(x,\nu),

where \nabla is the Levi-Civita connection of (Σ,g)(\Sigma,g). Applying the Newton transformation Tk1T_{k-1} on both sides of (15), we get

(16) (nk+1)(nk1)ΣV0Hk1Hkg¯(x,ν)dA\displaystyle(n-k+1)\binom{n}{k-1}\int_{\Sigma}V_{0}H_{k-1}-H_{k}\overline{g}(x,\nu)dA
=\displaystyle= ΣdivΣ(Tk1(xT))𝑑A\displaystyle\int_{\Sigma}\mbox{div}_{\Sigma}(T_{k-1}(x^{T}))dA
=\displaystyle= ΣTk1(xT,μ)𝑑s\displaystyle\int_{\partial\Sigma}T_{k-1}(x^{T},\mu)ds
=\displaystyle= cosθΣSk1;μg¯(x,ν¯)𝑑s.\displaystyle\cos\theta\int_{\partial\Sigma}S_{k-1;\mu}\overline{g}(x,\overline{\nu})ds.

In the last equality, we use (6) and the fact that μ\mu is a principal direction of Σ\Sigma and g¯(x,N¯)=0\overline{g}(x,\overline{N})=0. Therefore, Tk1(xT,μ)=Sk1;μg¯(x,μ)=cosθSk1;μg¯(x,ν¯)T_{k-1}(x^{T},\mu)=S_{k-1;\mu}\overline{g}(x,\mu)=\cos\theta S_{k-1;\mu}\overline{g}(x,\overline{\nu}) where Sk1;μS_{k-1;\mu} is given by

Sk1;μ=1α1<<αk1n1μ{eα1,,eαk1}λα1λα2λαk1.\displaystyle S_{k-1;\mu}=\sum_{\begin{subarray}{c}1\leq\alpha_{1}<\dots<\alpha_{k-1}\leq n-1\\ \mu\notin\{e_{\alpha_{1}},\dots,e_{\alpha_{k-1}}\}\end{subarray}}\lambda_{\alpha_{1}}\lambda_{\alpha_{2}}\cdots\lambda_{\alpha_{k-1}}.

On the other hand, let Un+1=g¯(ν,euEn+1)xg¯(x,ν)euEn+1U_{n+1}=\overline{g}(\nu,e^{-u}E_{n+1})x-\overline{g}(x,\nu)e^{-u}E_{n+1} be a tangent vector field on Σ\Sigma. We can easily see that g¯(Un+1,ν)=0\overline{g}(U_{n+1},\nu)=0 along Σ\Sigma. Then using Proposition 1, we have

(17) g¯(eiUn+1,ej)=\displaystyle\overline{g}(\nabla_{e_{i}}U_{n+1},e_{j})= g¯(¯ei(g¯(ν,euEn+1)xTg¯(x,ν)(euEn+1)T),ej)\displaystyle\overline{g}\left(\overline{\nabla}_{e_{i}}\left(\overline{g}(\nu,e^{-u}E_{n+1})x^{T}-\overline{g}(x,\nu)(e^{-u}E_{n+1})^{T}\right),e_{j}\right)
=\displaystyle= eug¯(x,ν)g¯(euEn+1,ei)g¯(x,ej)+g¯(euEn+1,ν)(V0δij\displaystyle-e^{-u}\overline{g}(x,\nu)\overline{g}(e^{-u}E_{n+1},e_{i})\overline{g}(x,e_{j})+\overline{g}(e^{-u}E_{n+1},\nu)(V_{0}\delta_{ij}
hijg¯(x,ν))g¯(x,ν)[eu(δijg¯(x,euEn+1)\displaystyle-h_{ij}\overline{g}(x,\nu))-\overline{g}(x,\nu)\big{[}e^{-u}(\delta_{ij}\overline{g}(x,e^{-u}E_{n+1})
g¯(euEn+1,ei)g¯(x,ej)hijg¯(ν,euEn+1)]\displaystyle-\overline{g}(e^{-u}E_{n+1},e_{i})\overline{g}(x,e_{j})-h_{ij}\overline{g}(\nu,e^{-u}E_{n+1})\big{]}
=\displaystyle= (g¯(euEn+1,ν)V0eug¯(x,ν)g¯(x,euEn+1))δij\displaystyle\left(\overline{g}(e^{-u}E_{n+1},\nu)V_{0}-e^{-u}\overline{g}(x,\nu)\overline{g}(x,e^{-u}E_{n+1})\right)\delta_{ij}
=\displaystyle= g¯(Yn+1,ν)δij,\displaystyle\overline{g}(Y_{n+1},\nu)\delta_{ij},

where we use the fact that euV0En+1x,En+1x=Yn+1e^{-u}V_{0}E_{n+1}-\langle x,E_{n+1}\rangle x=Y_{n+1}. Again, applying the Newton transformation Tk1T_{k-1} on both sides of (17) and integrating on Σ\Sigma, we have

(18) (nk+1)(nk1)ΣHk1g¯(Yn+1,ν)𝑑A\displaystyle(n-k+1)\binom{n}{k-1}\int_{\Sigma}H_{k-1}\overline{g}(Y_{n+1},\nu)dA
=\displaystyle= Σ(Tk1)ijg¯(eiUn+1,ej)𝑑A=ΣTk1(Un+1,μ)𝑑s\displaystyle\int_{\Sigma}(T_{k-1})_{ij}\overline{g}(\nabla_{e_{i}}U_{n+1},e_{j})dA=\int_{\partial\Sigma}T_{k-1}(U_{n+1},\mu)ds
=\displaystyle= ΣSk1;μg¯(Un+1,μ)𝑑s\displaystyle\int_{\partial\Sigma}S_{k-1;\mu}\overline{g}(U_{n+1},\mu)ds
=\displaystyle= ΣSk1;μg¯(x,ν¯)𝑑s,\displaystyle\int_{\partial\Sigma}S_{k-1;\mu}\overline{g}(x,\overline{\nu})ds,

where in the last equality, we use (6) on Σ\partial\Sigma to obtain

g¯(Un+1,μ)=\displaystyle\overline{g}(U_{n+1},\mu)= g¯(ν,euEn+1)g¯(x,μ)g¯(x,ν)g¯(euEn+1,μ)\displaystyle\overline{g}(\nu,e^{-u}E_{n+1})\overline{g}(x,\mu)-\overline{g}(x,\nu)\overline{g}(e^{-u}E_{n+1},\mu)
=\displaystyle= (cos2θg¯(x,ν¯)+sin2θg¯(x,ν¯))=g¯(x,ν¯).\displaystyle\left(\cos^{2}\theta\overline{g}(x,\overline{\nu})+\sin^{2}\theta\overline{g}(x,\overline{\nu})\right)=\overline{g}(x,\overline{\nu}).

Combining (16) and (18), we obtain the Minkowski type formula (14).

3.2. A Minkowski type formula for capillary hypersurfaces in Poincaré half space model

In Poincaré half space model, the inner normal vector field along LϕL_{\phi} is given by

(19) N¯=xn+1(cosϕEn+1+sinϕa).\overline{N}=x_{n+1}(-\cos\phi E_{n+1}+\sin\phi a).

Similarly, it can be easily to observe that the conformal Killing vector field Xn+1=x~En+1X_{n+1}=\widetilde{x}-E_{n+1} is tangential to LϕL_{\phi}. Indeed, we have

(20) g¯(Xn+1,N¯)=\displaystyle\overline{g}(X_{n+1},\overline{N})= xn+1g¯(x~En+1,cosϕEn+1+sinϕa)\displaystyle x_{n+1}\overline{g}(\widetilde{x}-E_{n+1},-\cos\phi E_{n+1}+\sin\phi a)
=\displaystyle= cosϕ+sinϕxaxn+1+cosϕxn+1\displaystyle-\cos\phi+\sin\phi\frac{x_{a}}{x_{n+1}}+\frac{\cos\phi}{x_{n+1}}
=\displaystyle= 0,\displaystyle 0,

where in the last equality we use (2).

xn+1x_{n+1}xax_{a}ϕ\phiBintB^{int}Σ\displaystyle{\color[rgb]{0.72,0.91,0.53}\definecolor[named]{pgfstrokecolor}{rgb}{0.72,0.91,0.53}\Sigma}N¯\overline{N}LϕL_{\phi}ν¯\overline{\nu}ν\nuμ\mu
Figure 3. Capillary hypersurface Σ\Sigma supported on an equidistant hypersurface LϕL_{\phi}.

Thus, we have the following Minkowski type formula analogously.

Proposition 4.

Let x:ΣBϕ,aintx:\Sigma\subset B^{\mathrm{int}}_{\phi,a} be a compact, immersed capillary hypersurface supported on Lϕ,aL_{\phi,a}, intersecting Lϕ,aL_{\phi,a} at a constant angle θ\theta. For k{1,2,,n}k\in\{1,2,\dots,n\}, we have

(21) Σ[Hk1(Vn+1cosθsecϕg¯(x~,ν))Hkg¯(Xn+1,ν)]𝑑A=0.\int_{\Sigma}[H_{k-1}(V_{n+1}-\cos\theta\sec\phi\overline{g}(\widetilde{x},\nu))-H_{k}\overline{g}(X_{n+1},\nu)]dA=0.
Proof.

As the proof of Proposition 3, restricting (11) to Σ\Sigma and using the divergence theorem, we have

(22) (nk+1)(nk1)ΣVn+1Hk1Hkg¯(Xn+1,ν)dA\displaystyle(n-k+1)\binom{n}{k-1}\int_{\Sigma}V_{n+1}H_{k-1}-H_{k}\overline{g}(X_{n+1},\nu)dA
=\displaystyle= ΣTk1(Xn+1,μ)𝑑s\displaystyle\int_{\partial\Sigma}T_{k-1}(X_{n+1},\mu)ds
=\displaystyle= cosθΣSk1;μg¯(Xn+1,ν¯)𝑑s,\displaystyle\cos\theta\int_{\partial\Sigma}S_{k-1;\mu}\overline{g}(X_{n+1},\overline{\nu})ds,

where in the last equality we use (6), (20) and the fact that μ\mu is a principal direction of Σ\Sigma along Σ\partial\Sigma.

On the other hand, consider a vector field Zn+1=g¯(x,ν)E¯n+1g¯(E¯n+1,ν)xZ_{n+1}=\overline{g}(x,\nu)\overline{E}_{n+1}-\overline{g}(\overline{E}_{n+1},\nu)x. By a direct computation, we have

g¯(eiZn+1,ej)\displaystyle\overline{g}(\nabla_{e_{i}}Z_{n+1},e_{j})
=\displaystyle= g¯(ei,E¯n+1)g¯(x~,ν)g¯(E¯n+1,ej)+g¯(x~,ei)g¯(E¯n+1,ν)g¯(E¯n+1,ej)\displaystyle-\overline{g}(e_{i},\overline{E}_{n+1})\overline{g}(\widetilde{x},\nu)\overline{g}(\overline{E}_{n+1},e_{j})+\overline{g}(\widetilde{x},e_{i})\overline{g}(\overline{E}_{n+1},\nu)\overline{g}(\overline{E}_{n+1},e_{j})
+hikg¯(x~,ek)g¯(E¯n+1,ej)+g¯(x~,ν)(g¯(E¯n+1,ei)g¯(E¯n+1,ej)δij)\displaystyle+h_{ik}\overline{g}(\widetilde{x},e_{k})\overline{g}(\overline{E}_{n+1},e_{j})+\overline{g}(\widetilde{x},\nu)(\overline{g}(\overline{E}_{n+1},e_{i})\overline{g}(\overline{E}_{n+1},e_{j})-\delta_{ij})
g¯(E¯n+1,ei)g¯(E¯n+1,ν)g¯(x~,ej)hikg¯(E¯n+1,ek)g¯(x~,ej)\displaystyle-\overline{g}(\overline{E}_{n+1},e_{i})\overline{g}(\overline{E}_{n+1},\nu)\overline{g}(\widetilde{x},e_{j})-h_{ik}\overline{g}(\overline{E}_{n+1},e_{k})\overline{g}(\widetilde{x},e_{j})
g¯(E¯n+1,ν)[g¯(ei,E¯n+1)g¯(x~,ej)+g¯(x~,ej)g¯(E¯n+1,ei)]\displaystyle-\overline{g}(\overline{E}_{n+1},\nu)\left[-\overline{g}(e_{i},\overline{E}_{n+1})\overline{g}(\widetilde{x},e_{j})+\overline{g}(\widetilde{x},e_{j})\overline{g}(\overline{E}_{n+1},e_{i})\right]
=\displaystyle= δijg¯(x~,ν)+g¯(E¯n+1,ν)(g¯(x~,ei)g¯(E¯n+1,ej)g¯(E¯n+1,ei)g¯(x~,ej)).\displaystyle-\delta_{ij}\overline{g}(\widetilde{x},\nu)+\overline{g}(\overline{E}_{n+1},\nu)\left(\overline{g}(\widetilde{x},e_{i})\overline{g}(\overline{E}_{n+1},e_{j})-\overline{g}(\overline{E}_{n+1},e_{i})\overline{g}(\widetilde{x},e_{j})\right).

Therefore,

(23) 12(g¯(eiZn+1,ej)+g¯(ejZn+1,ei))=g¯(x~,ν)δij.\displaystyle\frac{1}{2}\left(\overline{g}(\nabla_{e_{i}}Z_{n+1},e_{j})+\overline{g}(\nabla_{e_{j}}Z_{n+1},e_{i})\right)=-\overline{g}(\widetilde{x},\nu)\delta_{ij}.

Using the divergence Theorem, (6), (7) and (19), we have

(24) (nk+1)(nk1)ΣHk1g¯(x~,ν)𝑑A\displaystyle-(n-k+1)\binom{n}{k-1}\int_{\Sigma}H_{k-1}\overline{g}(\widetilde{x},\nu)dA
=\displaystyle= ΣTk1(Zn+1,μ)𝑑s\displaystyle\int_{\partial\Sigma}T_{k-1}(Z_{n+1},\mu)ds
=\displaystyle= ΣSk1;μg¯(Zn+1,μ)𝑑s\displaystyle\int_{\partial\Sigma}S_{k-1;\mu}\overline{g}(Z_{n+1},\mu)ds
=\displaystyle= ΣSk1;μ[g¯(x~,cosθN¯+sinθν¯)g¯(E¯n+1,sinθN¯+cosθν¯)\displaystyle\int_{\partial\Sigma}S_{k-1;\mu}\left[\overline{g}(\widetilde{x},-\cos\theta\overline{N}+\sin\theta\overline{\nu})\overline{g}(\overline{E}_{n+1},\sin\theta\overline{N}+\cos\theta\overline{\nu})\right.
g¯(E¯n+1,cosθN¯+sinθν¯)g¯(x~,sinθN¯+cosθν¯)]ds\displaystyle-\left.\overline{g}(\overline{E}_{n+1},-\cos\theta\overline{N}+\sin\theta\overline{\nu})\overline{g}(\widetilde{x},\sin\theta\overline{N}+\cos\theta\overline{\nu})\right]ds
=\displaystyle= ΣSk1;μ[(cosϕ+sinϕxaxn+1)g¯(E¯n+1,ν¯)+cosϕg¯(x~,ν¯)]𝑑s\displaystyle-\int_{\partial\Sigma}S_{k-1;\mu}\left[(-\cos\phi+\sin\phi\frac{x_{a}}{x_{n+1}})\overline{g}(\overline{E}_{n+1},\overline{\nu})+\cos\phi\overline{g}(\widetilde{x},\overline{\nu})\right]ds
=\displaystyle= ΣcosϕSk1;μg¯(Xn+1,ν¯)𝑑s,\displaystyle-\int_{\partial\Sigma}\cos\phi S_{k-1;\mu}\overline{g}(X_{n+1},\overline{\nu})ds,

where the last equality we use the fact from (2) that on ΣLϕ,a\partial\Sigma\subset L_{\phi,a}, xn+1=tanϕxa+1x_{n+1}=\tan\phi x_{a}+1. As consequence, the proof follows by combining (22) and (24). ∎

4. Proofs of Theorem 1 and Theorem 3

Reilly established an important integral formula for general Riemannian manifolds in [20]. Reilly’s formula is very powerful to obtain some properties on the hypersurfaces in Riemannian manifolds with non-negative Ricci curvatures. For instance, Ros used it to establish a Heintze-Karcher type inequality in [21]. Later on, Li and Xia [13] generalized it to a weighted Reilly-type formula to obtain a broader range of results. Let (¯,Δ¯,¯2)(\overline{\nabla},\overline{\Delta},\overline{\nabla}^{2}) and (,Δ,2)(\nabla,\Delta,\nabla^{2}) be the triples of Levi-Civita connection, Laplacian and Hessian operator on (Ω¯,g¯)(\overline{\Omega},\overline{g}) and (Ω,g)(\partial\Omega,g) respectively. Then their Reilly-type formula is as follows:

Theorem 7 (See [13]).

Let (Ω¯,g¯)(\overline{\Omega},\overline{g}) be a compact Riemannian manifold with piecewise smooth boundary Ω\partial\Omega, VC(Ω¯)V\in C^{\infty}(\overline{\Omega}) be a positive smooth function. Suppose ¯2VV\frac{\overline{\nabla}^{2}V}{V} is continuous up to Ω\partial\Omega. Then for any smooth function fC(Ω¯)f\in C^{\infty}(\overline{\Omega}), the following identity holds.

(25) ΩV((Δ¯fΔ¯VVf)2|¯2f¯2VVf|2)𝑑Ω\displaystyle\int_{\Omega}V\left(\left(\overline{\Delta}f-\frac{\overline{\Delta}V}{V}f\right)^{2}-\left|\overline{\nabla}^{2}f-\frac{\overline{\nabla}^{2}V}{V}f\right|^{2}\right)d\Omega
=\displaystyle= Ω(Δ¯Vg¯¯2V+VRic¯)(¯f¯VVf,¯f¯VVf)𝑑Ω\displaystyle\int_{\Omega}\left(\overline{\Delta}V\overline{g}-\overline{\nabla}^{2}V+V\overline{\mbox{Ric}}\right)\left(\overline{\nabla}f-\frac{\overline{\nabla}V}{V}f,\overline{\nabla}f-\frac{\overline{\nabla}V}{V}f\right)d\Omega
+ΩV(fνΩVνΩVf)(ΔfΔVVf)𝑑A\displaystyle+\int_{\partial\Omega}V\left(f_{\nu_{\partial\Omega}}-\frac{V_{\nu_{\partial\Omega}}}{V}f\right)\left(\Delta f-\frac{\Delta V}{V}f\right)dA
ΩVg((fνΩVνΩVf),fVVf)𝑑A\displaystyle-\int_{\partial\Omega}Vg\left(\nabla\left(f_{\nu_{\partial\Omega}}-\frac{V_{\nu_{\partial\Omega}}}{V}f\right),\nabla f-\frac{\nabla V}{V}f\right)dA
+ΩnVH1(fνΩVνΩVf)2𝑑A\displaystyle+\int_{\partial\Omega}nVH_{1}\left(f_{\nu_{\partial\Omega}}-\frac{V_{\nu_{\partial\Omega}}}{V}f\right)^{2}dA
+Ω(hVνΩVg)(fVVf,fVVf)𝑑A,\displaystyle+\int_{\partial\Omega}\left(h-\frac{V_{\nu_{\partial\Omega}}}{V}g\right)\left(\nabla f-\frac{\nabla V}{V}f,\nabla f-\frac{\nabla V}{V}f\right)dA,

where νΩ\nu_{\partial\Omega} is the outer normal vector field on Ω\partial\Omega, h(,)h(\cdot,\cdot) and H1H_{1} are the second fundamental form and mean curvature on Ω\partial\Omega respectively, and Ric\mathrm{Ric} is the Ricci curvature in Ω\Omega.

Let Ω\Omega be a domain in 𝔹+n+1\mathbb{B}_{+}^{n+1} enclosed by Σ\Sigma and TT, where TT is a domain on PP enclosed by Σ\partial\Sigma. Inspired by the idea in [10], we will give a proof of Theorem 1 now.

Proof of Theorem 1.

For θ=π2\theta=\frac{\pi}{2}, we refer to [18]. Therefore we will deal with the case of θ(0,π2)\theta\in(0,\frac{\pi}{2}). We consider the following mixed boundary problem,

(29) {Δ¯f(n+1)f=1, in Ω;f=0, on Σ;fN¯=c0, on T,\displaystyle\left\{\begin{array}[]{ll}\overline{\Delta}f-(n+1)f=1,&\mbox{ in }\Omega;\\ f=0,&\mbox{ on }\Sigma;\\ f_{\overline{N}}=c_{0},&\mbox{ on }T,\end{array}\right.

where c0=nn+1cotθTV0𝑑ATΣV0𝑑sc_{0}=-\frac{n}{n+1}\cot\theta\frac{\int_{T}V_{0}dA_{T}}{\int_{\partial\Sigma}V_{0}ds}. Since the idea about the existence and regularity is well established in [10], we will leave details of the following conclusion in the appendix: The solution ff of (29) satisfies fC(Ω¯Σ)C2(ΩT)C1,α(Ω¯)f\in C^{\infty}(\overline{\Omega}\setminus\partial\Sigma)\cup C^{2}(\Omega\cup T)\cup C^{1,\alpha}(\overline{\Omega}) and |¯2f|g¯L1(T)|\overline{\nabla}^{2}f|_{\overline{g}}\in L^{1}(T). This makes ff applicable in the generalized Reilly’s formula (25) and by letting V=V0V=V_{0}. Using (25), we have

(30) nn+1ΩV0𝑑Ω=\displaystyle\frac{n}{n+1}\int_{\Omega}V_{0}d\Omega= nn+1ΩV0(Δ¯fΔ¯V0V0f)2𝑑Ω\displaystyle\frac{n}{n+1}\int_{\Omega}V_{0}\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{0}}{V_{0}}f\right)^{2}d\Omega
\displaystyle\geq ΩV0((Δ¯fΔ¯V0V0f)2|¯2f¯2V0V0f|2)𝑑Ω\displaystyle\int_{\Omega}V_{0}\left(\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{0}}{V_{0}}f\right)^{2}-\left|\overline{\nabla}^{2}f-\frac{\overline{\nabla}^{2}V_{0}}{V_{0}}f\right|^{2}\right)d\Omega
=\displaystyle= ΣnV0H1fν2𝑑A+c0T(V0ΔffΔV0)𝑑AT\displaystyle\int_{\Sigma}nV_{0}H_{1}f_{\nu}^{2}dA+c_{0}\int_{T}(V_{0}\Delta f-f\Delta V_{0})dA_{T}
=\displaystyle= ΣnV0H1fν2𝑑A+c0ΣV0g(Tf,ν¯)𝑑s.\displaystyle\int_{\Sigma}nV_{0}H_{1}f_{\nu}^{2}dA+c_{0}\int_{\partial\Sigma}V_{0}g(\nabla^{T}f,\overline{\nu})ds.

In the second equality we use the fact that ¯2V0=V0g¯\overline{\nabla}^{2}V_{0}=V_{0}\overline{g}, Ric¯=ng¯\overline{\mbox{Ric}}=n\overline{g} and (V0)N¯=g¯(x~,N¯)=0(V_{0})_{\overline{N}}=\overline{g}(\widetilde{x},\overline{N})=0 on TT, and in the third equality we apply the Green’s formula (See [17]).

Furthermore, since fC1(Ω¯)f\in C^{1}(\overline{\Omega}) and f=0f=0 on Σ\Sigma, we have fμ=0f_{\mu}=0. From the relation (6), we have the following:

ν¯=\displaystyle\overline{\nu}= cosθμ+sinθν\displaystyle\cos\theta\mu+\sin\theta\nu
=\displaystyle= cosθμ+sinθ(cosθN¯+sinθν¯),\displaystyle\cos\theta\mu+\sin\theta(-\cos\theta\overline{N}+\sin\theta\overline{\nu}),
=\displaystyle= cosθμsinθcosθN¯+sin2θν¯.\displaystyle\cos\theta\mu-\sin\theta\cos\theta\overline{N}+\sin^{2}\theta\overline{\nu}.

Since ν¯=1cosθμtanθN¯\overline{\nu}=\frac{1}{\cos\theta}\mu-\tan\theta\overline{N}, we have

c0ΣV0fν¯𝑑s=\displaystyle c_{0}\int_{\partial\Sigma}V_{0}f_{\overline{\nu}}ds= c0ΣV0g¯(¯ffN¯N¯,ν¯)𝑑s\displaystyle c_{0}\int_{\partial\Sigma}V_{0}\overline{g}(\overline{\nabla}f-f_{\overline{N}}\overline{N},\overline{\nu})ds
=\displaystyle= c0ΣV0g¯(Tf,1cosθμtanθN¯)𝑑s\displaystyle c_{0}\int_{\partial\Sigma}V_{0}\overline{g}(\nabla^{T}f,\frac{1}{\cos\theta}\mu-\tan\theta\overline{N})ds
=\displaystyle= c02tanθΣV0𝑑s\displaystyle-c_{0}^{2}\tan\theta\int_{\partial\Sigma}V_{0}ds
=\displaystyle= (nn+1)2cotθ(TV0𝑑AT)2ΣV0𝑑s.\displaystyle-\left(\frac{n}{n+1}\right)^{2}\cot\theta\frac{\left(\int_{T}V_{0}dA_{T}\right)^{2}}{\int_{\partial\Sigma}V_{0}ds}.

Therefore, (30) is equivalent to the following inequality,

(31) nn+1(ΩV0𝑑Ω+nn+1cotθ(TV0𝑑AT)2ΣV0𝑑s)ΣnV0H1fν2𝑑A.\displaystyle\frac{n}{n+1}\left(\int_{\Omega}V_{0}d\Omega+\frac{n}{n+1}\cot\theta\frac{\left(\int_{T}V_{0}dA_{T}\right)^{2}}{\int_{\partial\Sigma}V_{0}ds}\right)\geq\int_{\Sigma}nV_{0}H_{1}f_{\nu}^{2}dA.

On the other hand, by Green’s formula to the equation above, we have

ΩV0𝑑Ω=\displaystyle\int_{\Omega}V_{0}d\Omega= ΩV0(Δ¯fΔ¯V0V0f)𝑑Ω\displaystyle\int_{\Omega}V_{0}\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{0}}{V_{0}}f\right)d\Omega
=\displaystyle= Σ(V0fνf(V0)ν)𝑑A+T(V0fN¯f(V0)N¯)𝑑AT\displaystyle\int_{\Sigma}(V_{0}f_{\nu}-f(V_{0})_{\nu})dA+\int_{T}(V_{0}f_{\overline{N}}-f(V_{0})_{\overline{N}})dA_{T}
=\displaystyle= ΣV0fν𝑑A+c0TV0𝑑AT\displaystyle\int_{\Sigma}V_{0}f_{\nu}dA+c_{0}\int_{T}V_{0}dA_{T}
=\displaystyle= ΣV0fν𝑑Ann+1cotθ(TV0𝑑AT)2ΣV0𝑑s.\displaystyle\int_{\Sigma}V_{0}f_{\nu}dA-\frac{n}{n+1}\cot\theta\frac{\left(\int_{T}V_{0}dA_{T}\right)^{2}}{\int_{\partial\Sigma}V_{0}ds}.

Applying Hölder’s inequality, we obtain

(32) ΩV0𝑑Ω+nn+1cotθ(TV0𝑑νT)2ΣV0𝑑s\displaystyle\int_{\Omega}V_{0}d\Omega+\frac{n}{n+1}\cot\theta\frac{\left(\int_{T}V_{0}d\nu_{T}\right)^{2}}{\int_{\partial\Sigma}V_{0}ds}
=\displaystyle= ΣV0fμ𝑑A(ΣV0nH1𝑑A)12(ΣnV0H1fν2𝑑A)12.\displaystyle\int_{\Sigma}V_{0}f_{\mu}dA\leq\left(\int_{\Sigma}\frac{V_{0}}{nH_{1}}dA\right)^{\frac{1}{2}}\left(\int_{\Sigma}nV_{0}H_{1}f_{\nu}^{2}dA\right)^{\frac{1}{2}}.

Combining (31) and (32), we have

(33) ΣV0H1𝑑A(n+1)ΩV0𝑑Ω+ncotθ(TV0𝑑AT)2ΣV0𝑑s.\displaystyle\int_{\Sigma}\frac{V_{0}}{H_{1}}dA\geq(n+1)\int_{\Omega}V_{0}d\Omega+n\cot\theta\frac{\left(\int_{T}V_{0}dA_{T}\right)^{2}}{\int_{\partial\Sigma}V_{0}ds}.

If the equality in (33) holds, all the inequalities above become equalities. In particular from (30) we have,

¯2ffg¯=¯2f¯2V0V0f=1n+1(Δ¯fΔ¯V0V0f)g¯=1n+1g¯,\displaystyle\overline{\nabla}^{2}f-f\overline{g}=\overline{\nabla}^{2}f-\frac{\overline{\nabla}^{2}V_{0}}{V_{0}}f=\frac{1}{n+1}\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{0}}{V_{0}}f\right)\overline{g}=\frac{1}{n+1}\overline{g},

which is equivalent to

(34) ¯2(f+1n+1)=(f+1n+1)g¯\displaystyle\overline{\nabla}^{2}\left(f+\frac{1}{n+1}\right)=\left(f+\frac{1}{n+1}\right)\overline{g}

Restricting (34) to Σ\Sigma where f=0f=0, we can see that

fνhij=1n+1g¯ij.f_{\nu}h_{ij}=\frac{1}{n+1}\overline{g}_{ij}.

This implies that Σ\Sigma is totally umbilical. ∎

Remark.

In the case of θ=π2\theta=\frac{\pi}{2}, we note that the last term of the right hand side in (30) vanishes, therefore there is no need for the Green’s formula used in the second equality in (30), the theory in [14] is sufficient. See the Appendix for detail.

Before giving a proof of Theorem 3, we need the following identities.

Proposition 5.

On the embedded capillary hypersurface Σ𝔹+n+1\Sigma\subset\mathbb{B}^{n+1}_{+} supported on PP, the following identities holds,

(35) TV0𝑑AT=Σg¯(Yn+1,ν)𝑑A,\displaystyle\int_{T}V_{0}dA_{T}=\int_{\Sigma}\overline{g}(Y_{n+1},\nu)dA,

and

(36) sinθΣV0𝑑s=ΣnH1g¯(Yn+1,ν)𝑑A.\displaystyle\sin\theta\int_{\partial\Sigma}V_{0}ds=\int_{\Sigma}nH_{1}\overline{g}(Y_{n+1},\nu)dA.
Proof.

Since Yn+1Y_{n+1} is a Killing vector field, then div¯(Yn+1)=0\overline{\mathrm{div}}(Y_{n+1})=0 on Ω\Omega. Integrate on Ω\Omega, we have

0=\displaystyle 0= Ωdiv¯(Yn+1)𝑑Ω\displaystyle\int_{\Omega}\overline{\mathrm{div}}(Y_{n+1})d\Omega
=\displaystyle= Σg¯(Yn+1,ν)𝑑A+Tg¯(Yn+1,N¯)𝑑AT.\displaystyle\int_{\Sigma}\overline{g}(Y_{n+1},\nu)dA+\int_{T}\overline{g}(Y_{n+1},\overline{N})dA_{T}.

Since g¯(Yn+1,N¯)=V0\overline{g}(Y_{n+1},\overline{N})=-V_{0}, we have

TV0𝑑AT=Σg¯(Yn+1,ν)𝑑A.\displaystyle\int_{T}V_{0}dA_{T}=\int_{\Sigma}\overline{g}(Y_{n+1},\nu)dA.

Restricting (10) to Σ\Sigma and using the divergence theorem, we have

ΣnH1g¯(Yn+1,ν)𝑑A=ΣdivΣ(PΣYn+1)𝑑A=Σg¯(Yn+1,μ)𝑑s=sinθΣV0𝑑s,\displaystyle-\int_{\Sigma}nH_{1}\overline{g}(Y_{n+1},\nu)dA=\int_{\Sigma}\mathrm{div}_{\Sigma}(P_{\Sigma}Y_{n+1})dA=\int_{\partial\Sigma}\overline{g}(Y_{n+1},\mu)ds=-\sin\theta\int_{\partial\Sigma}V_{0}ds,

where PΣP_{\Sigma} denotes the projection on TΣT\Sigma. ∎

We are now ready to prove Theorem 3.

Proof of the Theorem 3.

Since Σ\Sigma is compact, there exists a t(0,π)t\in(0,\pi) such that Σ\Sigma is contained in a domain enclosed by PP and consider 1-parameter family fo equidistant hypersurfaces StS_{t}.

St={xn+1:|x+tantEn+1|2=1cos2t,t(0,π2)}.\displaystyle S_{t}=\left\{x\in\mathbb{H}^{n+1}:|x+\tan tE_{n+1}|^{2}=\frac{1}{\cos^{2}t},\;t\in(0,\frac{\pi}{2})\right\}.

When t0+t\rightarrow 0+, StS_{t} tends to n+1\partial_{\infty}\mathbb{H}^{n+1}. On the other hand, when tπ2t\rightarrow\frac{\pi}{2}-, StS_{t} tends to PP. Given that Σ\partial\Sigma is compact, and St=Pn+1\partial S_{t}=\partial P\subset\partial_{\infty}\mathbb{H}^{n+1}, therefore when tπ2t\rightarrow\frac{\pi}{2}-, StS_{t} would only touch the interior of Σ\Sigma other than its boundary. As tπ2t\rightarrow\frac{\pi}{2}, there must exist a t0t_{0} such that St0S_{t_{0}} touches int(Σ)\mbox{int}(\Sigma) at first and tangent to Σ\Sigma at a point pint(Σp\in\mbox{int}(\Sigma). Hence, we have h(p)λS(p)>0h(p)\geq\lambda_{S}(p)>0. Then H1H1(p)>0H_{1}\equiv H_{1}(p)>0. See the Figure 4.

Σ\displaystyle{\color[rgb]{0.72,0.91,0.53}\definecolor[named]{pgfstrokecolor}{rgb}{0.72,0.91,0.53}\Sigma}StS_{t}P\displaystyle P
Figure 4. Existence of the convex point on capillary hypersurface Σ\Sigma supported on PP

Now applying (35) and (36) to the inequality (1), we can see that (1) is equivalent to the following:

(37) ΣV0H1𝑑A(n+1)ΩV0𝑑Ω+cosθ(Σg¯(Yn+1,ν)𝑑A)2ΣH1g¯(Yn+1,ν)𝑑A.\displaystyle\int_{\Sigma}\frac{V_{0}}{H_{1}}dA\geq(n+1)\int_{\Omega}V_{0}d\Omega+\cos\theta\frac{\left(\int_{\Sigma}\overline{g}(Y_{n+1},\nu)dA\right)^{2}}{\int_{\Sigma}H_{1}\overline{g}(Y_{n+1},\nu)dA}.

Since H1H_{1} is constant, (37) can be written as

(38) ΣV0cosθg¯(Yn+1,ν)H1𝑑A\displaystyle\int_{\Sigma}\frac{V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu)}{H_{1}}dA\geq (n+1)ΩV0𝑑Ω\displaystyle(n+1)\int_{\Omega}V_{0}d\Omega
=\displaystyle= Ωdiv¯(x)𝑑Ω\displaystyle\int_{\Omega}\overline{\mathrm{div}}(x)d\Omega
=\displaystyle= Σg¯(x,ν)𝑑A.\displaystyle\int_{\Sigma}\overline{g}(x,\nu)dA.

The last equality holds because g¯(x,N¯)=0\overline{g}(x,\overline{N})=0 on TT. Since H1H_{1} is constant, from Proposition 3 (k=1k=1) we can see that the equality in (38) holds. According to the condition in Theorem 1, we obtain that Σ\Sigma is umbilical, but it is not totally geodesic since H1>0H_{1}>0. ∎

Remark.

We consider the equality case of Theorem 1. From [25, Remark 4.2], we know that any function in {VC(n+1):¯2V=Vg¯}\{V\in C^{\infty}(\mathbb{H}^{n+1}):\overline{\nabla}^{2}V=V\overline{g}\} can be represented by the linear combination of the n+2n+2 functions {V0,V1,,Vn+1}\{V_{0},V_{1},\dots,V_{n+1}\}, where Vi=2x,Ei1|x|2V_{i}=\frac{2\langle x,E_{i}\rangle}{1-|x|^{2}}. From (34) we can see that there exists constants B,ciB,\,c_{i}\in\mathbb{R} and a constant vector (in the Euclidean sense) an+1a\in\mathbb{R}^{n+1} such that

f+1n+1=\displaystyle f+\frac{1}{n+1}= BV0+i=1n+1ciVi\displaystyle BV_{0}+\sum\limits_{i=1}^{n+1}c_{i}V_{i}
=\displaystyle= B1+|x|21|x|2+2x,C1|x|2\displaystyle B\frac{1+|x|^{2}}{1-|x|^{2}}+\frac{2\langle x,C\rangle}{1-|x|^{2}}
=\displaystyle= B(1+|x|2)+2x,C1|x|2,\displaystyle\frac{B(1+|x|^{2})+2\langle x,C\rangle}{1-|x|^{2}},

where C=i=1n+1ciEiC=\sum\limits_{i=1}^{n+1}c_{i}E_{i}.

  1. (1)

    When B1n+1B\neq-\frac{1}{n+1}, Σ\Sigma lies in a hypersurface defined by

    {xn+1:|x+(n+1)1+(n+1)BC|2=1(n+1)B1+(n+1)B+((n+1)|C|1+(n+1)B)2}.\displaystyle\left\{x\in\mathbb{H}^{n+1}:\left|x+\frac{(n+1)}{1+(n+1)B}C\right|^{2}=\frac{1-(n+1)B}{1+(n+1)B}+\left(\frac{(n+1)|C|}{1+(n+1)B}\right)^{2}\right\}.
  2. (2)

    When B=1n+1B=-\frac{1}{n+1} and C0C\neq 0, Σ\Sigma lies in a hypersurface defined by

    {xn+1:x,a=1/|C|}.\displaystyle\left\{x\in\mathbb{H}^{n+1}:\langle x,a\rangle=1/|C|\right\}.

Therefore, Σ\Sigma is a totally umbilical hypersurface in n+1\mathbb{H}^{n+1} and can be a geodesic ball, a horosphere or an equidistant hypersurfaces, and B,C,an+1B,\,C\in\mathbb{R},a\in\mathbb{R}^{n+1} can be chosen such that Σ\Sigma is a part of one of the illustrated above and compact (compactness excludes the totally geodesic case). For example, a horosphere can be seen Figure 5.

a\displaystyle aEOFP\displaystyle PΣ\displaystyle\Sigma
Figure 5. Compact capillary horosphere

From the Figure 5, we can see

|OF|=|1(n+1)B1+(n+1)B|,|OE|=|(n+1)1+(n+1)BC|.\displaystyle|OF|=\sqrt{\left|\frac{1-(n+1)B}{1+(n+1)B}\right|},\quad|OE|=\left|\frac{(n+1)}{1+(n+1)B}C\right|.

5. Proofs of Theorem 2 and Theorem 4

In this section, we consider the case when the supporting hypersurface is either an equidistant hypersurface or a horospheres. Before we give a proof of Theorem 2, we need the following lemma.

Proposition 6.

On the embedded capillary hypersurfaces Σ+n+1\Sigma\hookrightarrow\mathbb{R}^{n+1}_{+} supported on Lϕ,aL_{\phi,a} in the Poincaré half space model, the following identities holds

(39) Σg¯(x~,ν)𝑑A=cosϕTVn+1𝑑AT;\int_{\Sigma}\overline{g}(\widetilde{x},\nu)dA=\cos\phi\int_{T}V_{n+1}dA_{T};
(40) Σg¯(a,ν)𝑑A=sinϕTVn+1𝑑AT;\int_{\Sigma}\overline{g}(a,\nu)dA=-\sin\phi\int_{T}V_{n+1}dA_{T};
(41) Σ(nH1g¯(x~,ν))𝑑A=Σg¯(x~,μ)𝑑s;\int_{\Sigma}(-nH_{1}\overline{g}(\widetilde{x},\nu))dA=\int_{\partial\Sigma}\overline{g}(\widetilde{x},\mu)ds;
(42) Σ(nH1g¯(a,ν))𝑑A=Σg¯(a,μ)𝑑s;\int_{\Sigma}(-nH_{1}\overline{g}(a,\nu))dA=\int_{\partial\Sigma}\overline{g}(a,\mu)ds;

and

(43) Σncosθg¯(x~,ν)𝑑A=Σcosϕg¯(Xn+1,μ)𝑑s.\int_{\Sigma}n\cos\theta\overline{g}(\widetilde{x},\nu)dA=\int_{\partial\Sigma}\cos\phi\overline{g}(X_{n+1},\mu)ds.
Proof.

To prove (39) and (41), we know from (13) that div¯x~=0\overline{\mathrm{div}}\widetilde{x}=0. Let Ω\Omega be a domain enclosed by Σ\Sigma and TT. Using the divergence Theorem, we have

0=\displaystyle 0= Ωdiv¯x~𝑑Ω\displaystyle\int_{\Omega}\overline{\mathrm{div}}\widetilde{x}d\Omega
=\displaystyle= Σg¯(x~,ν)𝑑A+Tg¯(N¯,x~)𝑑AT\displaystyle\int_{\Sigma}\overline{g}(\widetilde{x},\nu)dA+\int_{T}\overline{g}(\overline{N},\widetilde{x})dA_{T}
=\displaystyle= Σg¯(x~,ν)𝑑A+T1xn+1(cosϕxn+1+sinϕxa)𝑑AT\displaystyle\int_{\Sigma}\overline{g}(\widetilde{x},\nu)dA+\int_{T}\frac{1}{x_{n+1}}(-\cos\phi x_{n+1}+\sin\phi x_{a})dA_{T}
=\displaystyle= Σg¯(x~,ν)𝑑ATcosϕVn+1dAT.\displaystyle\int_{\Sigma}\overline{g}(\widetilde{x},\nu)dA-\int_{T}\cos\phi V_{n+1}dA_{T}.

In the last equality, we use (2).

On the other hand, restricting (13) on Σ\Sigma, we have

Σ(nH1g¯(x~,ν))𝑑A=ΣdivΣ(PΣx~)𝑑A=Σg¯(x~,μ)𝑑s,\displaystyle\int_{\Sigma}\left(-nH_{1}\overline{g}(\widetilde{x},\nu)\right)dA=\int_{\Sigma}\mathrm{div}_{\Sigma}(P_{\Sigma}\widetilde{x})dA=\int_{\partial\Sigma}\overline{g}(\widetilde{x},\mu)ds,

where PΣP_{\Sigma} denotes the projection on TΣT\Sigma. Then we obtain (39) and (41). Using (12), we can prove (42) and (40). Moreover, (43) can be noted from in the proof of Proposition 4. ∎

Proof of Theorem 2.

Consider the following mixed boundary problem in the Poincaré half space model,

(44) {Δ¯f(n+1)f=1, in Ω;f=0 on Σ;fN¯cosϕf=c0 on T,\left\{\begin{array}[]{ll}\overline{\Delta}f-(n+1)f=1,&\mbox{ in }\Omega;\\ f=0&\mbox{ on }\Sigma;\\ f_{\overline{N}}-\cos\phi f=c_{0}&\mbox{ on }T,\end{array}\right.

where

c0=nn+1cosθTVn+1𝑑ATΣg¯(cosϕx~+sinϕa,μ)𝑑s.c_{0}=-\frac{n}{n+1}\cos\theta\frac{\int_{T}V_{n+1}dA_{T}}{\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\phi a,\mu)ds}.

According to the Lieberman’s theory (see the appendix), there exists a solution ff such that fC(Ω¯Σ)C2(ΩT)C1,α(Ω¯)f\in C^{\infty}(\overline{\Omega}\setminus\partial\Sigma)\cup C^{2}(\Omega\cup T)\cup C^{1,\alpha}(\overline{\Omega}) and |¯2f|g¯L1(T)|\overline{\nabla}^{2}f|_{\overline{g}}\in L^{1}(T). Therefore, for the same reason as the proof of Theorem 1, we can insert the solution ff to the generalized Reilly’s type formula (25).

nn+1ΩVn+1𝑑Ω=\displaystyle\frac{n}{n+1}\int_{\Omega}V_{n+1}d\Omega= nn+1ΩVn+1(Δ¯fΔ¯Vn+1Vn+1f)2𝑑Ω\displaystyle\frac{n}{n+1}\int_{\Omega}V_{n+1}\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{n+1}}{V_{n+1}}f\right)^{2}d\Omega
\displaystyle\geq ΩVn+1((Δ¯fΔ¯Vn+1Vn+1f)2|¯2f¯2Vn+1Vn+1f|2)𝑑Ω\displaystyle\int_{\Omega}V_{n+1}\left(\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{n+1}}{V_{n+1}}f\right)^{2}-\left|\overline{\nabla}^{2}f-\frac{\overline{\nabla}^{2}V_{n+1}}{V_{n+1}}f\right|^{2}\right)d\Omega
=\displaystyle= ΣnVn+1H1fν2𝑑A+c0T(Vn+1ΔffΔVn+1)𝑑AT\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f_{\nu}^{2}dA+c_{0}\int_{T}(V_{n+1}\Delta f-f\Delta V_{n+1})dA_{T}
+nc02cosϕTVn+1𝑑AT\displaystyle+nc_{0}^{2}\cos\phi\int_{T}V_{n+1}dA_{T}
=\displaystyle= ΣnVn+1H1fν2𝑑A+c0ΣVn+1g¯(Tf,ν¯)𝑑s+nc02Σg¯(x~,ν)𝑑A\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f^{2}_{\nu}dA+c_{0}\int_{\partial\Sigma}V_{n+1}\overline{g}(\nabla^{T}f,\overline{\nu})ds+nc_{0}^{2}\int_{\Sigma}\overline{g}(\widetilde{x},\nu)dA
=\displaystyle= ΣnVn+1H1fν2𝑑A+c0ΣVn+1g¯(Tf,ν¯)𝑑s\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f^{2}_{\nu}dA+c_{0}\int_{\partial\Sigma}V_{n+1}\overline{g}(\nabla^{T}f,\overline{\nu})ds
+c02cosϕcosθΣg¯(Xn+1,μ)𝑑s,\displaystyle+\frac{c_{0}^{2}\cos\phi}{\cos\theta}\int_{\partial\Sigma}\overline{g}(X_{n+1},\mu)ds,

where we use the fact that (Vn+1)N¯=1xn+12En+1,N¯=cosϕVn+1(V_{n+1})_{\overline{N}}=-\frac{1}{x_{n+1}^{2}}\langle E_{n+1},\overline{N}\rangle=\cos\phi V_{n+1} and (43) in the last equality. On Σ\partial\Sigma, using (6) and the boundary condition in (44), we have

g¯(Tf,ν¯)=\displaystyle\overline{g}(\nabla^{T}f,\overline{\nu})= g¯(¯fg¯(¯f,N¯)N¯,1cosθμ)\displaystyle\overline{g}(\overline{\nabla}f-\overline{g}(\overline{\nabla}f,\overline{N})\overline{N},\frac{1}{\cos\theta}\mu)
=\displaystyle= c0cosθg¯(N¯,μ).\displaystyle-\frac{c_{0}}{\cos\theta}\overline{g}(\overline{N},\mu).

Then we have

nn+1ΩVn+1𝑑Ω\displaystyle\frac{n}{n+1}\int_{\Omega}V_{n+1}d\Omega\geq ΣnVn+1H1fν2𝑑Ac02cosθΣVn+1g¯(N¯,μ)𝑑s\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f_{\nu}^{2}dA-\frac{c_{0}^{2}}{\cos\theta}\int_{\partial\Sigma}V_{n+1}\overline{g}(\overline{N},\mu)ds
+c02cosϕcosθΣg¯(Xn+1,μ)𝑑s\displaystyle+\frac{c_{0}^{2}\cos\phi}{\cos\theta}\int_{\partial\Sigma}\overline{g}(X_{n+1},\mu)ds
=\displaystyle= ΣnVn+1H1fν2𝑑Ac02cosθΣg¯(cosϕx~+sinϕa,μ)𝑑s\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f_{\nu}^{2}dA-\frac{c_{0}^{2}}{\cos\theta}\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\phi a,\mu)ds
=\displaystyle= ΣnVn+1H1fν2𝑑A(nn+1)2cosθ(TVn+1𝑑AT)2Σg¯(cosϕx~+sinθa,μ)𝑑s.\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f_{\nu}^{2}dA-\left(\frac{n}{n+1}\right)^{2}\cos\theta\frac{\left(\int_{T}V_{n+1}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\theta a,\mu)ds}.

Therefore,

(45) nn+1(ΩVn+1𝑑Ω+nn+1cosθ(TVn+1𝑑AT)2Σg¯(cosϕx~+sinθa,μ)𝑑s)\displaystyle\frac{n}{n+1}\left(\int_{\Omega}V_{n+1}d\Omega+\frac{n}{n+1}\cos\theta\frac{\left(\int_{T}V_{n+1}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\theta a,\mu)ds}\right)
\displaystyle\geq ΣnVn+1H1fν2𝑑A.\displaystyle\int_{\Sigma}nV_{n+1}H_{1}f_{\nu}^{2}dA.

On the other hand, using the divergence theorem we have

(46) ΩVn+1𝑑Ω=\displaystyle\int_{\Omega}V_{n+1}d\Omega= ΩVn+1(Δ¯fΔ¯Vn+1Vn+1f)𝑑Ω\displaystyle\int_{\Omega}V_{n+1}\left(\overline{\Delta}f-\frac{\overline{\Delta}V_{n+1}}{V_{n+1}}f\right)d\Omega
=\displaystyle= Σ(Vn+1fνf(Vn+1)ν)𝑑A+T(Vn+1fN¯f(Vn+1)N¯)𝑑AT\displaystyle\int_{\Sigma}(V_{n+1}f_{\nu}-f(V_{n+1})_{\nu})dA+\int_{T}(V_{n+1}f_{\overline{N}}-f(V_{n+1})_{\overline{N}})dA_{T}
=\displaystyle= ΣVn+1fν𝑑A+c0TVn+1𝑑AT\displaystyle\int_{\Sigma}V_{n+1}f_{\nu}dA+c_{0}\int_{T}V_{n+1}dA_{T}
=\displaystyle= ΣVn+1fν𝑑Ann+1cosθ(TVn+1𝑑AT)2Σg¯(cosϕx~+sinϕa,μ)𝑑s.\displaystyle\int_{\Sigma}V_{n+1}f_{\nu}dA-\frac{n}{n+1}\cos\theta\frac{\left(\int_{T}V_{n+1}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\phi a,\mu)ds}.

Combining (45) and (46) and using the similar argument as the proof of Theorem 1, we prove the inequality in Theorem 2. The equality case is also the same as Theorem 1. ∎

At this point, we are ready to prove the Alexandrov type theorem for the capillary hypersurface in Bϕ,aintB^{\mathrm{int}}_{\phi,a}.

Proof of the Theorem 4.

In order to prove the existence of convex points on Σ\Sigma, we consider both the case that ϕ>0\phi>0 and the case that ϕ=0\phi=0 independently.
Case 1 (ϕ>0\phi>0): Now Lϕ,aL_{\phi,a} is an equidistant hypersurface. Let η(0,ϕ)\eta\in(0,\phi), which will be determined later. We consider a set

Pη={x~n+1:xn+1=tanη(xa+1)}.P_{\eta}=\{\widetilde{x}\in\mathbb{H}^{n+1}:x_{n+1}=\tan\eta(x_{a}+1)\}.

Let pPηp\in P_{\eta} and SpS_{p} be a family of equidistant hypersurface defined as

Sp={x~n+1:|x~p|=|pdist(p,Pη)| and pPη},S_{p}=\{\widetilde{x}\in\mathbb{H}^{n+1}:|\widetilde{x}-p|=|p-\mathrm{dist}_{\mathbb{R}}(p,\partial P_{\eta})|\mbox{ and }p\in P_{\eta}\},

where dist\mathrm{dist}_{\mathbb{R}} is the Euclidean distance in the Poincaré half space model. Since Σ\Sigma is a compact capillary hypersurface supported on Lϕ,aL_{\phi,a}, we can find a proper p0Pηp_{0}\in P_{\eta} sufficiently far (in the Euclidean sense) from Pη\partial P_{\eta}, such that Σ\Sigma is contained in the domain enclosed by Sp0S_{p_{0}} and Lϕ,aL_{\phi,a} and η>0\eta>0 guarantees that Σ\Sigma stays in the interior side of Sp0S_{p_{0}}. Then for any pPp\in P the contact angle θ0\theta_{0} between Lϕ,aL_{\phi,a} and Sp0S_{p_{0}} is constant satisfying θ0=π2ϕ+η\theta_{0}=\frac{\pi}{2}-\phi+\eta.

Since θ+ϕ>π2\theta+\phi>\frac{\pi}{2}, there exists an η(0,ϕ)\eta\in(0,\phi) such that θ>θ0=π2ϕ+η\theta>\theta_{0}=\frac{\pi}{2}-\phi+\eta.

Now moving p0p_{0} toward Pη\partial P_{\eta} such that dist(p0,Pη)\mathrm{dist}_{\mathbb{R}}(p_{0},\partial P_{\eta}) goes to 0, we can see that there exists p1p_{1} such that Sp1S_{p_{1}} firstly touches Σ\Sigma at some qint(Σ)q\in\mathrm{int}(\Sigma) since θ>θ0\theta>\theta_{0}. Meanwhile, since η>0\eta>0, Σ\Sigma remains within the interior side of Sp1S_{p_{1}}. See Figure 6.

xn+1x_{n+1}xax_{a}ϕ\phiBintB^{int}Σ\displaystyle{\color[rgb]{0.72,0.91,0.53}\definecolor[named]{pgfstrokecolor}{rgb}{0.72,0.91,0.53}\Sigma}η\etaN¯\overline{N}θ0\theta_{0}PηP_{\eta}LϕL_{\phi}p0p_{0}Sp0S_{p_{0}}
Figure 6. Existence of the convex point for hypersurfaces supported on a equidistant hypersurface

Then we can see that the principal curvature λ(q)λSp1(q)>0\lambda(q)\geq\lambda_{S_{p_{1}}}(q)>0 of Σ\Sigma, then H1H1(q)>0H_{1}\equiv H_{1}(q)>0.
Case 2 (ϕ=0\phi=0): Now Lϕ,aL_{\phi,a} is a horosphere and θ=π2\theta=\frac{\pi}{2} by assumption. Consider a set

Pr0={xn+1:xn+1=r0},P_{r_{0}}=\{x\in\mathbb{H}^{n+1}:x_{n+1}=r_{0}\},

for 0<r0<10<r_{0}<1. Let SR,pS_{R,p} be the hypersurface defined by

SR,p={xn+1:|x~p|=R,pPr0}.S_{R,p}=\{x\in\mathbb{H}^{n+1}:|\widetilde{x}-p|=R,p\in P_{r_{0}}\}.

Since Σ\Sigma is compact, there must exist p0Pr0p_{0}\in P_{r_{0}} and a positive R0R_{0} big enough such that Σ\Sigma is contained in the domain enclosed by SR0,p0S_{R_{0},p_{0}} and Lϕ,aL_{\phi,a}. Now fix p0p_{0} and shrinking the radius R0R_{0} of SR0,p0S_{R_{0},p_{0}} in the Euclidean sense until it touches Σ\Sigma at pp for the first time when the radius is R1R_{1}. Since the contact angle of SR1,p0S_{R_{1},p_{0}} is less than π2\frac{\pi}{2}, the firstly touching point qq must be the interior point on Σ\Sigma. Since the contact angle of SR1,p0S_{R_{1},p_{0}} and n+1:={xn+1=0}\partial_{\infty}\mathbb{H}^{n+1}:=\{x_{n+1}=0\} is greater than π2\frac{\pi}{2}, Σ\Sigma is staying in the interior side of SR1,p0S_{R_{1},p_{0}}. See Figure 7.

L0L_{0}Pr0P_{r_{0}}{}Sp0,R0S_{p_{0},R_{0}}Σ\Sigmap0p_{0}xn+1x_{n+1}xax_{a}
Figure 7. Existence of the convex point for hypersurfaces supported on a horosphere

Then it holds that λ(q)λSR1,p0>0\lambda(q)\geq\lambda_{S_{R_{1},p_{0}}}>0 and H1H1(q)>0H_{1}\equiv H_{1}(q)>0. Therefore the Heintze-Karcher inequality in Theorem 2 can be applied. And from (39), (39), (42) and (40), we are able to rewrite (3) as

ΣVn+1H1𝑑A\displaystyle\int_{\Sigma}\frac{V_{n+1}}{H_{1}}dA\geq (n+1)ΩVn+1𝑑Ω+ncosθ(TVn+1𝑑AT)2Σg¯(cosϕx~+sinϕa,μ)𝑑s\displaystyle(n+1)\int_{\Omega}V_{n+1}d\Omega+n\cos\theta\frac{\left(\int_{T}V_{n+1}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\overline{g}(-\cos\phi\widetilde{x}+\sin\phi a,\mu)ds}
=\displaystyle= (n+1)ΩVn+1𝑑Ω+cosθ(TVn+1𝑑AT)2H1Σg¯(cosϕx~sinϕa,ν)𝑑A\displaystyle(n+1)\int_{\Omega}V_{n+1}d\Omega+\cos\theta\frac{\left(\int_{T}V_{n+1}dA_{T}\right)^{2}}{H_{1}\int_{\Sigma}\overline{g}(\cos\phi\widetilde{x}-\sin\phi a,\nu)dA}
=\displaystyle= (n+1)ΩVn+1𝑑Ω+secθcosϕH1Σg¯(x~,ν)𝑑A.\displaystyle(n+1)\int_{\Omega}V_{n+1}d\Omega+\frac{\sec\theta\cos\phi}{H_{1}}\int_{\Sigma}\overline{g}(\widetilde{x},\nu)dA.

From (11), we have

(n+1)Vn+1dΩ=\displaystyle(n+1)V_{n+1}d\Omega= Ωdiv¯Xn+1𝑑Ω\displaystyle\int_{\Omega}\overline{\mathrm{div}}X_{n+1}d\Omega
=\displaystyle= Σg¯(Xn+1,ν)𝑑A+Tg¯(Xn+1,N¯)𝑑AT\displaystyle\int_{\Sigma}\overline{g}(X_{n+1},\nu)dA+\int_{T}\overline{g}(X_{n+1},\overline{N})dA_{T}
=\displaystyle= Σg¯(Xn+1,ν)𝑑A.\displaystyle\int_{\Sigma}\overline{g}(X_{n+1},\nu)dA.

Now the rest is the same as the prove of Theorem 3. Together with Minkowski type formula (21), we prove that Σ\Sigma is umbilical. ∎

Remark.

Indeed, the following example indicate that there exists a totally umbilical capillary hypersurface in Bϕ,aintB^{\mathrm{int}}_{\phi,a} without any convex point. Consider a totally geodesic hypersurface is given as follows

Σ={x~n+1:δ(x~,x~)=2},\Sigma=\{\widetilde{x}\in\mathbb{H}_{n+1}:\delta(\widetilde{x},\widetilde{x})=2\},

and the supporting hypersurface is Lπ6L_{\frac{\pi}{6}}. Then the totally geodesic Σ\Sigma is a capillary hypersurface supported on Lπ6L_{\frac{\pi}{6}} and the contact angle θ=arccos640.912\theta=\arccos\frac{\sqrt{6}}{4}\approx 0.912, but it is apparent to have not any convex point. Also we can see that θ+ϕ1.435<π2\theta+\phi\approx 1.435<\frac{\pi}{2}, which is not satisfying the condition in Theorem 4. See Figure 8.

Refer to caption
Figure 8. Capillary hypersurfaces in Bϕ,aintB^{\mathrm{int}}_{\phi,a} without any convex point

We can see from the proof of Theorem 4 that the assumption “ϕ+θ>π2\phi+\theta>\frac{\pi}{2} for ϕ>0\phi>0 or θ=π2\theta=\frac{\pi}{2} for ϕ=0\phi=0” can be replaced by H1>0H_{1}>0. Then, we have the following corollary.

Corollary 1.

Let Σ\Sigma be a compactly embedded CMC capillary hypersurface contained in Bϕ,aintB^{\mathrm{int}}_{\phi,a} supported on Lϕ,aL_{\phi,a}, and the contact angle satisfies θ(0,π2]\theta\in(0,\frac{\pi}{2}]. Assume H1>0H_{1}>0, then Σ\Sigma is umbilical except for being totally geodesic.

6. Capillary hypersurfaces in a geodesic ball

In this section, we give an Alexandrov type theorem for capillary hypersurfaces in a geodesic ball. Consider a geodesic ball BRB_{R} centered at the origin in the Poincaré ball model, where RR the hyperbolic radius. Then BRB_{R} can be given by

BR={xn+1:g(x,x)R2}={xn+1:|x|2Rδ2},B_{R}=\{x\in\mathbb{H}^{n+1}:g(x,x)\leq R^{2}\}=\{x\in\mathbb{H}^{n+1}:|x|^{2}\leq R_{\delta}^{2}\},

where RδR_{\delta} is the Euclidean radius of BRB_{R}. Thus, the following may be easily noted:

coshR=1+Rδ21Rδ2 and sinhR=2Rδ1Rδ2.\cosh R=\frac{1+R_{\delta}^{2}}{1-R_{\delta}^{2}}\quad\mbox{ and }\quad\sinh R=\frac{2R_{\delta}}{1-R_{\delta}^{2}}.

Then the unit outward normal N¯\overline{N} of BRB_{R} satisfies that

(47) N¯=1sinhRx.\overline{N}=\frac{1}{\sinh R}x.

In the paper of Wang and Xia (See [25]), the authors give a family of conformal Killing vector fields XaX_{a}. It is given by

Xa=21Rδ2[δ(x,a)x12(|x|2+Rδ2)a],X_{a}=\frac{2}{1-R_{\delta}^{2}}\left[\delta(x,a)x-\frac{1}{2}(|x|^{2}+R_{\delta}^{2})a\right],

where aa is a constant vector with respect to the Euclidean metric (𝔹n+1,δ)(\mathbb{B}^{n+1},\delta) satisfying that

12(g¯(¯E¯AXa,E¯B)+g¯(¯E¯BXa,E¯A))=Vag¯,\frac{1}{2}\left(\overline{g}(\overline{\nabla}_{\overline{E}_{A}}X_{a},\overline{E}_{B})+\overline{g}(\overline{\nabla}_{\overline{E}_{B}}X_{a},\overline{E}_{A})\right)=V_{a}\overline{g},

where Va=2δ(x,a)1|x|2V_{a}=\frac{2\delta(x,a)}{1-|x|^{2}} satisfying ¯2Va=Vag¯\overline{\nabla}^{2}V_{a}=V_{a}\overline{g}. A direct computation shows that

(48) coshRXa=Vaxsinh2RYa.\cosh RX_{a}=V_{a}x-\sinh^{2}RY_{a}.
BR\displaystyle B_{R}O\displaystyle ON¯\overline{N}μ\muν\nuν¯\overline{\nu}Σ\Sigma
Figure 9. Capillary hypersurface Σ\Sigma supported on a geodesic ball BRB_{R}

We have the following properties analogously,

Proposition 7.

Let Σ\Sigma be an embedded capillary hypersurface supported on the geodesic ball BRB_{R}. Let Ω\Omega be the domain enclosed by Σ\Sigma and BR\partial B_{R} and Ω=ΣT\partial\Omega=\Sigma\cup T, we have

(49) TVa𝑑AT=sinhRΣg¯(Ya,ν)𝑑A,\int_{T}V_{a}dA_{T}=-\sinh R\int_{\Sigma}\overline{g}(Y_{a},\nu)dA,
(50) Σg¯(Ya,μ)𝑑s=ΣnH1g¯(Ya,ν)𝑑A,\int_{\partial\Sigma}\overline{g}(Y_{a},\mu)ds=-\int_{\Sigma}nH_{1}\overline{g}(Y_{a},\nu)dA,
(51) Σ(Va+cosθsinhRg¯(Ya,ν)H1g¯(Xa,ν))𝑑A=0,\int_{\Sigma}\left(V_{a}+\cos\theta\sinh R\overline{g}(Y_{a},\nu)-H_{1}\overline{g}(X_{a},\nu)\right)dA=0,

and

(52) ΣnsinhRcosθg¯(Ya,ν)𝑑A=Σg¯(Xa,μ)𝑑s.\int_{\Sigma}n\sinh R\cos\theta\overline{g}(Y_{a},\nu)dA=\int_{\partial\Sigma}\overline{g}(X_{a},\mu)ds.
Sketch of the proof.

The proof of (49) and (50) can be seen easily from the proof of other supporting hypersurfaces cases, using the fact that YaY_{a} is a Killing vector field, and (51) and (52) are given in the proof of [25, Prop 4.4]. ∎

Consider the following mixed boundary problem

(53) {Δ¯f(n+1)f=1 in Ω;f=0 on Σ;fN¯cothRf=c0 on T.\left\{\begin{array}[]{ll}\overline{\Delta}f-(n+1)f=1&\mbox{ in }\Omega;\\ f=0&\mbox{ on }\Sigma;\\ f_{\overline{N}}-\coth Rf=c_{0}&\mbox{ on }T.\end{array}\right.

Where c0=nn+1cosθTVa𝑑ATΣsinhRg¯(Ya,μ)𝑑sc_{0}=-\frac{n}{n+1}\cos\theta\frac{\int_{T}V_{a}dA_{T}}{\int_{\partial\Sigma}\sinh R\overline{g}(Y_{a},\mu)ds} is chosen. The existence and regularity of the solution to the problem (53) can be obtain from Lieberman’s theory. See the Appendix.

Theorem 8.

Let BR,a+={xBR:Va>0}B_{R,a+}=\{x\in B_{R}:V_{a}>0\} be a half geodesic ball where VaV_{a} is positive. Let ΣBR,a+\Sigma\subset B_{R,a+} be a compact, embedded capillary hypersurface in BR,a+B_{R,a+}, and the supporting hypersurface is BR\partial B_{R}. Let Ω\Omega be the domain enclosed by Σ\Sigma and Lϕ,aL_{\phi,a}. If the contact angle θ(0,π2]\theta\in(0,\frac{\pi}{2}], and the mean curvature H1>0H_{1}>0 on Σ\Sigma, we have

(54) ΣVaH1𝑑A(n+1)ΩVa𝑑Ω+ncosθ(TVa𝑑AT)2ΣsinhRg¯(Ya,μ)𝑑s,\displaystyle\int_{\Sigma}\frac{V_{a}}{H_{1}}dA\geq(n+1)\int_{\Omega}V_{a}d\Omega+n\cos\theta\frac{\left(\int_{T}V_{a}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\sinh R\overline{g}(Y_{a},\mu)ds},

where μ\mu is the outward unit conormal vector field of Σ\partial\Sigma in Σ\Sigma. In addition, the equality holds if and only if Σ\Sigma is umbilical.

Proof.

Let V=VaV=V_{a} and ff be the solution of (53) in (25). Using the relations (47), (48) and Proposition 7 and noting that (Va)N¯=cothRVa(V_{a})_{\overline{N}}=\coth RV_{a}, we have

nn+1ΩVa𝑑Ω\displaystyle\frac{n}{n+1}\int_{\Omega}V_{a}d\Omega\geq ΣnVaH1fν2𝑑A+c0T(VaffVa)𝑑AT+c02TncothRVadAT\displaystyle\int_{\Sigma}nV_{a}H_{1}f_{\nu}^{2}dA+c_{0}\int_{T}(V_{a}\triangle f-f\triangle V_{a})dA_{T}+c_{0}^{2}\int_{T}n\coth RV_{a}dA_{T}
=\displaystyle= ΣVaH1fν2𝑑A+c0ΣVafν¯𝑑s+c02TncothRVadAT\displaystyle\int_{\Sigma}V_{a}H_{1}f_{\nu}^{2}dA+c_{0}\int_{\partial\Sigma}V_{a}f_{\overline{\nu}}ds+c_{0}^{2}\int_{T}n\coth RV_{a}dA_{T}
=\displaystyle= ΣnVaH1fν2𝑑Ac02cosθΣVag¯(μ,1sinhRx)𝑑s\displaystyle\int_{\Sigma}nV_{a}H_{1}f_{\nu}^{2}dA-\frac{c_{0}^{2}}{\cos\theta}\int_{\partial\Sigma}V_{a}\overline{g}(\mu,\frac{1}{\sinh R}x)ds
c02coshRΣng¯(Ya,ν)𝑑A\displaystyle-c_{0}^{2}\cosh R\int_{\Sigma}n\overline{g}(Y_{a},\nu)dA
=\displaystyle= ΣnVaH1fν2𝑑Ac02cosθΣg¯(μ,cothRXa+sinhRYa)𝑑s\displaystyle\int_{\Sigma}nV_{a}H_{1}f_{\nu}^{2}dA-\frac{c_{0}^{2}}{\cos\theta}\int_{\partial\Sigma}\overline{g}(\mu,\coth RX_{a}+\sinh RY_{a})ds
+c02cosθΣcothRg¯(Ya,μ)𝑑s\displaystyle+\frac{c_{0}^{2}}{\cos\theta}\int_{\partial\Sigma}\coth R\overline{g}(Y_{a},\mu)ds
=\displaystyle= ΣnVaH1fν2𝑑Ac02cosθΣsinhRg¯(μ,Ya)𝑑s.\displaystyle\int_{\Sigma}nV_{a}H_{1}f_{\nu}^{2}dA-\frac{c_{0}^{2}}{\cos\theta}\int_{\partial\Sigma}\sinh R\overline{g}(\mu,Y_{a})ds.

Therefore,

(55) nn+1(ΩVa𝑑Ω+nn+1cosθ(TVa𝑑AT)2ΣsinhRg¯(Ya,μ)𝑑s)ΣnVaH1fν2𝑑A.\frac{n}{n+1}\left(\int_{\Omega}V_{a}d\Omega+\frac{n}{n+1}\cos\theta\frac{\left(\int_{T}V_{a}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\sinh R\overline{g}(Y_{a},\mu)ds}\right)\geq\int_{\Sigma}nV_{a}H_{1}f_{\nu}^{2}dA.

On the other hand,

ΩVa𝑑Ω=Ω(Va¯ff¯Va)𝑑Ω=ΣVafν𝑑A+c0TVa𝑑AT.\displaystyle\int_{\Omega}V_{a}d\Omega=\int_{\Omega}(V_{a}\overline{\triangle}f-f\overline{\triangle}V_{a})d\Omega=\int_{\Sigma}V_{a}f_{\nu}dA+c_{0}\int_{T}V_{a}dA_{T}.

We have

(56) ΣVafν𝑑A=\displaystyle\int_{\Sigma}V_{a}f_{\nu}dA= ΩVa𝑑Ωc0TVa𝑑AT\displaystyle\int_{\Omega}V_{a}d\Omega-c_{0}\int_{T}V_{a}dA_{T}
=\displaystyle= ΩVa𝑑Ω+nn+1cosθ(TVa𝑑AT)2ΣsinhRg¯(Ya,μ)𝑑s.\displaystyle\int_{\Omega}V_{a}d\Omega+\frac{n}{n+1}\cos\theta\frac{\left(\int_{T}V_{a}dA_{T}\right)^{2}}{\int_{\partial\Sigma}\sinh R\overline{g}(Y_{a},\mu)ds}.

Hence, using (55) and (56) and the similar argument as the proof of Theorem 1, we prove the inequality in Theorem 8. The equality case is also the same as Theorem 1. ∎

Using (49), (50), (51) in Proposition 7 and the fact that g¯(Xa,N¯)=0\overline{g}(X_{a},\overline{N})=0, we can prove the following Alexandrov type theorem.

Theorem 9.

Let Σ\Sigma be an embedded CMC capillary hypersurface contained in BR,a+B_{R,a+} supported on BR\partial B_{R}. If θ(0,π2]\theta\in(0,\frac{\pi}{2}] and H1>0H_{1}>0, then Σ\Sigma is umbilical except for being totally geodesic.

Remark.

Since we do not have the existence of convex points on Σ\Sigma, we assume that H1>0H_{1}>0.

7. Other rigidity results for capillary hypersurfaces

In this section, we will prove the Theorem 5 and Theorem 6. The proof of Theorem 5 is due to the Minkowski type formula in Proposition 3.

The proof of Theorem 5.

Since (Hk/Hl)(p)=α>0(H_{k}/H_{l})(p)=\alpha>0, from Newton-Maclaurin inequality, we have

Hk1HlHkHl1=αHlHl1.\displaystyle H_{k-1}H_{l}\geq H_{k}H_{l-1}=\alpha H_{l}H_{l-1}.

and since Hl>0H_{l}>0 are positive on Σ\Sigma, we have Hk1αHl10H_{k-1}-\alpha H_{l-1}\geq 0. On the other hand, applying Proposition 3, we have

(57) 0=\displaystyle 0= Σ[Hk1(V0cosθg¯(Yn+1,ν))Hkg¯(x,ν)]𝑑A\displaystyle\int_{\Sigma}\left[H_{k-1}(V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu))-H_{k}\overline{g}(x,\nu)\right]dA
=\displaystyle= Σ[Hk1(V0cosθg¯(Yn+1,ν))αHlg¯(x,ν)]𝑑A\displaystyle\int_{\Sigma}\left[H_{k-1}(V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu))-\alpha H_{l}\overline{g}(x,\nu)\right]dA
=\displaystyle= Σ(Hk1αHl1)(V0cosθg¯(Yn+1,ν))𝑑A.\displaystyle\int_{\Sigma}(H_{k-1}-\alpha H_{l-1})(V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu))dA.

Now, we consider the term V0cosθg¯(Yn+1,ν)V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu). We see that

g¯(Yn+1,Yn+1)=\displaystyle\overline{g}(Y_{n+1},Y_{n+1})= 14(1+|x|2)2g¯(En+1,En+1)+x,En+12g¯(x,x)\displaystyle\frac{1}{4}(1+|x|^{2})^{2}\overline{g}(E_{n+1},E_{n+1})+\langle x,E_{n+1}\rangle^{2}\overline{g}(x,x)
x,En+1(1+|x|2)g¯(x,En+1)\displaystyle-\langle x,E_{n+1}\rangle(1+|x|^{2})\overline{g}(x,E_{n+1})
=\displaystyle= (1+|x|2)24x,En+12(1|x|2)2.\displaystyle\frac{(1+|x|^{2})^{2}-4\langle x,E_{n+1}\rangle^{2}}{(1-|x|^{2})^{2}}.

Since x,En+1>0\langle x,E_{n+1}\rangle>0 on int(Σ)\mbox{int}(\Sigma), we have

|g¯(Yn+1,ν)|g¯(Yn+1,Yn+1)<V0,|\overline{g}(Y_{n+1},\nu)|\leq\sqrt{\overline{g}(Y_{n+1},Y_{n+1})}<V_{0},

and therefore

(58) V0cosθg¯(Yn+1,ν)>0.\displaystyle V_{0}-\cos\theta\overline{g}(Y_{n+1},\nu)>0.

Hence, combining (57) and (58), we have Hk1αHl1=0H_{k-1}-\alpha H_{l-1}=0. This results in

(59) Hk1Hl=HkHl1.\displaystyle H_{k-1}H_{l}=H_{k}H_{l-1}.

Hence, examining the equality condition of Newton-Maclaurin inequality, we obtain that Σ\Sigma is totally umbilical, completing the proof of Theorem 5. ∎

Now we give a proof of Theorem 6.

Proof of Theorem 6.

Let GG be a smooth function defined by G=nH1V0ng¯(x,ν)G=nH_{1}V_{0}-n\overline{g}(x,\nu). Integrating the Laplacian of GG, we have

ΣΔG𝑑A=\displaystyle\int_{\Sigma}\Delta GdA= ΣμGds\displaystyle\int_{\partial\Sigma}\nabla_{\mu}Gds
=\displaystyle= Σ(nH1nh(μ,μ))g¯(x,μ))ds\displaystyle\int_{\partial\Sigma}(nH_{1}-nh(\mu,\mu))\overline{g}(x,\mu))ds
=\displaystyle= cosθΣ(nH1nh(μ,μ))g¯(x,ν¯)𝑑s.\displaystyle\cos\theta\int_{\partial\Sigma}(nH_{1}-nh(\mu,\mu))\overline{g}(x,\overline{\nu})ds.

For any eT(Σ)e\in T(\partial\Sigma), we can see from(6) that

(60) h(e,e)=g¯(¯eν,e)=g¯(¯e(cosθN¯+sinθν¯),e)=sinθhΣ(e,e).\displaystyle h(e,e)=\overline{g}(\overline{\nabla}_{e}\nu,e)=\overline{g}(\overline{\nabla}_{e}(\cos\theta\overline{N}+\sin\theta\overline{\nu}),e)=\sin\theta h^{\partial\Sigma}(e,e).

By letting k=1k=1 in (18), we have

Σng¯(Yn+1,ν)𝑑A=Σg¯(x,ν¯)𝑑s.\displaystyle\int_{\Sigma}n\overline{g}(Y_{n+1},\nu)dA=\int_{\partial\Sigma}\overline{g}(x,\overline{\nu})ds.

Then we have

(61) ΣΔG𝑑A=\displaystyle\int_{\Sigma}\Delta GdA= cosθΣ(nH1nh(μ,μ))g¯(x,ν¯)𝑑s\displaystyle\cos\theta\int_{\partial\Sigma}(nH_{1}-nh(\mu,\mu))\overline{g}(x,\overline{\nu})ds
=\displaystyle= cosθΣ(n(nH1h(μ,μ))n(n1)H1)g¯(x,ν¯)𝑑s\displaystyle\cos\theta\int_{\partial\Sigma}(n(nH_{1}-h(\mu,\mu))-n(n-1)H_{1})\overline{g}(x,\overline{\nu})ds
=\displaystyle= cosθΣ(n(n1)sinθH1Σn(n1)H1)g¯(x,ν¯)𝑑s\displaystyle\cos\theta\int_{\partial\Sigma}(n(n-1)\sin\theta H^{\partial\Sigma}_{1}-n(n-1)H_{1})\overline{g}(x,\overline{\nu})ds
=\displaystyle= cosθ[Σn(n1)sinθH1Σg¯(x,ν¯)dsn2(n1)H1\displaystyle\cos\theta\left[\int_{\partial\Sigma}n(n-1)\sin\theta H^{\partial\Sigma}_{1}\overline{g}(x,\overline{\nu})ds-n^{2}(n-1)H_{1}\right.
×Σg¯(Yn+1,ν)dA]\displaystyle\quad\left.\times\int_{\Sigma}\overline{g}(Y_{n+1},\nu)dA\right]
=\displaystyle= cosθΣ(n(n1)sinθH1Σg¯(x,ν¯)n(n1)sinθV0)𝑑s\displaystyle\cos\theta\int_{\partial\Sigma}(n(n-1)\sin\theta H^{\partial\Sigma}_{1}\overline{g}(x,\overline{\nu})-n(n-1)\sin\theta V_{0})ds
=\displaystyle= 0.\displaystyle 0.

We use (36) in the fifth equality. In the last equality, we use a Minkowski type formula for closed hypersurface in PP (See [2]), which is the (n1)(n-1)-dimensional hyperbolic space.

On the other hand, it is well known that in hypersurface

ΔV0=nV0nH1¯νV0,\displaystyle\Delta V_{0}=nV_{0}-nH_{1}\overline{\nabla}_{\nu}V_{0},

and

Δg¯(x,ν)=nH1V0|h|2g¯(X,ν)n¯νV+ng¯(X,ν).\displaystyle\Delta\overline{g}(x,\nu)=nH_{1}V_{0}-|h|^{2}\overline{g}(X,\nu)-n\overline{\nabla}_{\nu}V+n\overline{g}(X,\nu).

Together with (61), we have

0=ΣΔG𝑑A=Σn(|h|2nH12)g¯(x,ν)𝑑A.\displaystyle 0=\int_{\Sigma}\Delta GdA=\int_{\Sigma}n(|h|^{2}-nH_{1}^{2})\overline{g}(x,\nu)dA.

Therefore, from the star-shapedness of Σ\Sigma, we can see that Σ\Sigma is umbilical. In addition, the totally geodesic case is excluded by the compactness assumption on Σ\Sigma. ∎

Remark.

All of our results require that Σ\Sigma is compact. Because Σ\Sigma is totally umbilical and contained in 𝔹+n+1\mathbb{B}^{n+1}_{+}, it can be compact and a part of a horosphere or an equidistant hypersurface simultaneously. If so, we notice that the contact angle θ\theta must lie in the open interval (0,π2)(0,\frac{\pi}{2}). See the Figure 10 and Figure 11.

n+1\displaystyle\mathbb{H}^{n+1}P\displaystyle PΣ\displaystyle\SigmaO\displaystyle O
Figure 10. Σ\Sigma is a horosphere
n+1\displaystyle\mathbb{H}^{n+1}P\displaystyle PΣ\displaystyle\SigmaO\displaystyle O
Figure 11. Σ\Sigma is an equidistant hypersurface
Remark.

Since the star-shapedness of Σ\Sigma implies that Σ\Sigma is an embedded hypersurface, Theorem 6 can be obtained immediately from Theorem 3 when the contact angle θ(0,π2]\theta\in(0,\frac{\pi}{2}].

Appendix

To obtain the existence and sufficient regularity of the solution of the mixed-boundary problem (29), (44) and (53), we write it in the conformal Euclidean space as

{f:=e2u(Δ¯δf+(n1)du(f))(n+1)f=1 in Ω;f=0 on Σ;fN¯γf=c0 on T,\displaystyle\left\{\begin{array}[]{ll}\mathcal{L}f:=e^{-2u}(\overline{\Delta}_{\delta}f+(n-1)du(f))-(n+1)f=1&\mbox{ in }\Omega;\\ f=0&\mbox{ on }\Sigma;\\ f_{\overline{N}}-\gamma f=c_{0}&\mbox{ on }T,\end{array}\right.

where ¯δ\overline{\nabla}_{\delta} and Δ¯δ\overline{\Delta}_{\delta} are the Levi-Civita connection and Laplacian with respect to the Euclidean metric δ\delta respectively. Since the conformal map preserves the angle, the condition of θ\theta is the same under both Euclidean metric and hyperbolic metric.

We notice that the coefficients aij=e2uδij,bi=(n1)e2uui,c=(n+1)a_{ij}=e^{-2u}\delta_{ij},\,b^{i}=(n-1)e^{-2u}u^{i},\;c=-(n+1) satisfy the conditions in [14], page 435, and γ0\gamma\geq 0 satisfies the condition in [23, Lemma 4.1]. From the theory of Lieberman and Theorem A.3 in [10], there exists a solution fC(Ω¯Σ)C2(int(Ω)T)f\in C^{\infty}(\overline{\Omega}\setminus\partial\Sigma)\cup C^{2}(\mathrm{int}(\Omega)\cup T) with respect to the conformal metric δ\delta.

Let dΣd_{\partial\Sigma} be the distance function from Σ\partial\Sigma and define Ωε={xΩ:dΣ(x)>ε}\Omega_{\varepsilon}=\{x\in\Omega:d_{\partial\Sigma}(x)>\varepsilon\}. To apply the result of Lieberman, we need to introduce the norm

|f|ab=supε>0εa+b|f|a,Ωε,\displaystyle|f|^{b}_{a}=\sup_{\varepsilon>0}\varepsilon^{a+b}|f|_{a,\Omega_{\varepsilon}},

where |f|a,Ωε=i=1ksupxΩε|δkf|+supx,yΩε|Dkf(x)Dkf(y)||xy|α|f|_{a,\Omega_{\varepsilon}}=\sum\limits_{i=1}^{k}\sup_{x\in\Omega_{\varepsilon}}|\nabla_{\delta}^{k}f|+\sup_{x,y\in\Omega_{\varepsilon}}\frac{|D^{k}f(x)-D^{k}f(y)|}{|x-y|^{\alpha}} for a>0a>0 and a=k+αa=k+\alpha for an integer kk and 0α<10\leq\alpha<1.

Let ff be the solution of the following boundary value problem

{f=g in Ω;f=0 on Σ;¯δf,N¯=h on T.\displaystyle\left\{\begin{array}[]{ll}\mathcal{L}f=g&\mbox{ in }\Omega;\\ f=0&\mbox{ on }\Sigma;\\ \langle\overline{\nabla}_{\delta}f,\overline{N}\rangle=h&\mbox{ on }T.\end{array}\right.

From the theory by Lieberman (See Theorem 4 in [15]) and Lemma A.1 in [10], we have the following estimate

|f|aλC(|g|a22λ+|h|a11λ+|g|0+|h|0).\displaystyle|f|_{a}^{-\lambda}\leq C(|g|_{a-2}^{2-\lambda}+|h|_{a-1}^{1-\lambda}+|g|_{0}+|h|_{0}).

Here we notice that the maximum principle (see [23, Lemma 4.1]) applied in [10, Lemma A.1] holds for general operator \mathcal{L} with c=(n+1)<0c=-(n+1)<0 and γ0\gamma\geq 0. Furthermore, from the proof of Theorem 4 in [15], the construction of the key Miller’s type barrier requires λ<π2θ\lambda<\frac{\pi}{2\theta} (See [15, Lemma 4.1]).

From the discussion by Jia-Wang-Xia (See [10], Page 8-9 and Lemma A.1), we can see that |¯δ2f|δL2(Ω)|\overline{\nabla}_{\delta}^{2}f|_{\delta}\in L^{2}(\Omega) requires |¯δ2f|δCdΣβ|\overline{\nabla}^{2}_{\delta}f|_{\delta}\leq Cd_{\partial\Sigma}^{-\beta} for β(0,1)\beta\in(0,1) and a>2a>2. Then, we can also obtain that |δ2f|δL1(T)|\nabla_{\delta}^{2}f|_{\delta}\in L^{1}(T).

Meanwhile in the proof of the Theorem 1, we need the regularity of ff satisfying fC1,α(Ω¯)f\in C^{1,\alpha}(\overline{\Omega}). From the definition and the monotonicity of the norm ||ab|\cdot|_{a}^{b}, we have |f|λ=|f|λλ|f|aλ|f|_{\lambda}=|f|_{\lambda}^{-\lambda}\leq|f|_{a}^{-\lambda}. Therefore it is sufficient for fC1,α(Ω¯)f\in C^{1,\alpha}(\overline{\Omega}) when λ>1\lambda>1.

In conclusion, all the condition for the applicable regularity will be satisfied if we let 1<λ<min{3,π2θ}1<\lambda<\mbox{min}\{3,\frac{\pi}{2\theta}\}, a=λ+32>2a=\frac{\lambda+3}{2}>2 and β=aλ=3λ2(0,1)\beta=a-\lambda=\frac{3-\lambda}{2}\in(0,1). Hence, the condition θ<π2\theta<\frac{\pi}{2} is sufficient. From a well-known fact, under conformal transformations of metrics we have

¯f=e2u¯δf,\displaystyle\overline{\nabla}f=e^{-2u}\overline{\nabla}_{\delta}f,

and

¯ij2f=¯δ,ij2fuifjujfiδ(¯δf,¯δu)δij.\displaystyle\overline{\nabla}^{2}_{ij}f=\overline{\nabla}^{2}_{\delta,\;ij}f-u_{i}f_{j}-u_{j}f_{i}-\delta(\overline{\nabla}_{\delta}f,\overline{\nabla}_{\delta}u)\delta_{ij}.

Then all the estimates above hold with respect to the hyperbolic metric g¯\overline{g}, that is, fC(Ω¯Σ)C2(ΩT)C1,α(Ω¯)f\in C^{\infty}(\overline{\Omega}\setminus\partial\Sigma)\cup C^{2}(\Omega\cup T)\cup C^{1,\alpha}(\overline{\Omega}) and |¯2f|g¯L1(T)|\overline{\nabla}^{2}f|_{\overline{g}}\in L^{1}(T). Now we obtain the existence and the regularity of the solution of (29).

References

  • [1] A. D. Aleksandrov “Uniqueness theorems for surfaces in the large. III” In Vestnik Leningrad. Univ. 13.7, 1958, pp. 14–26
  • [2] Simon Brendle “Constant mean curvature surfaces in warped product manifolds” In Publ. Math. Inst. Hautes Études Sci. 117, 2013, pp. 247–269
  • [3] Yimin Chen, Yingxiang Hu and Haizhong Li “Geometric inequalities for free boundary hypersurfaces in a ball” In Annals of Global Analysis and Geometry 62.1 Springer, 2022, pp. 33–45
  • [4] Ailana Fraser and Richard Schoen “Sharp eigenvalue bounds and minimal surfaces in the ball” In Invent. Math. 203.3, 2016, pp. 823–890
  • [5] Jinyu Guo, Guofang Wang and Chao Xia “Stable capillary hypersurfaces supported on a horosphere in the hyperbolic space” In Adv. Math. 409, Part A Elsevier, 2022, pp. 108641
  • [6] Jinyu Guo and Chao Xia “A partially overdetermined problem in domains with partial umbilical boundary in space forms” In Adv. Calc. Var. De Gruyter, 2022
  • [7] Yingxiang Hu, Yong Wei, Bo Yang and Tailong Zhou “A complete family of Alexandrov-Fenchel inequalities for convex capillary hypersurfaces in the half-space” In arXiv preprint arXiv:2209.12479, 2022
  • [8] Xiaohan Jia, Guofang Wang, Chao Xia and Xuwen Zhang “Heintze-Karcher inequality and capillary hypersurfaces in a wedge” In arXiv preprint arXiv:2209.13839, 2022
  • [9] Xiaohan Jia, Guofang Wang, Chao Xia and Xuwen Zhang “Alexandrov’s theorem for anisotropic capillary hypersurfaces in the half-space” In Arch. Ration. Mech. Anal. 247.2 Springer, 2023, pp. 25
  • [10] Xiaohan Jia, Chao Xia and Xuwen Zhang “A Heintze–Karcher-Type Inequality for Hypersurfaces with Capillary Boundary” In J. Geom. Anal. 33.6 Springer, 2023, pp. 177
  • [11] Sung-Eun Koh and Seung-Won Lee “Addendum to the paper: Sphere theorem by means of the ratio of mean curvature functions” In Glasg. Math. J. 43.2 Cambridge University Press, 2001, pp. 275–276
  • [12] Pierre Simon Laplace “Traité de mécanique céleste; suppléments au Livre X” In Gauthier-Villars, Œuvres Complètes 4, 1805
  • [13] Junfang Li and Chao Xia “An integral formula and its applications on sub-static manifolds” In J. Differential Geom. 113.3, 2019, pp. 493–518
  • [14] Gary M. Lieberman “Mixed boundary value problems for elliptic and parabolic differential equations of second order” In J. Math. Anal. Appl. 113.2, 1986, pp. 422–440
  • [15] Gary M.Lieberman “Optimal Hölder regularity for mixed boundary value problems” In J. Math. Anal. Appl. 143.2, 1989, pp. 572–586
  • [16] Sebastián Montiel “Unicity of constant mean curvature hypersurfaces in some Riemannian manifolds” In Indiana Univ. Math. J. 48.2, 1999, pp. 711–748
  • [17] Filomena Pacella and Giulio Tralli “Overdetermined problems and constant mean curvature surfaces in cones” In Rev. Mat. Iberoam. 36.3, 2020, pp. 841–867
  • [18] Juncheol Pyo “Rigidity theorems of hypersurfaces with free boundary in a wedge in a space form” In Pacific J. Math. 299.2, 2019, pp. 489–510
  • [19] Guohuan Qiu and Chao Xia “A generalization of Reilly’s formula and its applications to a new Heintze-Karcher type inequality” In Int. Math. Res. Not. IMRN, 2015, pp. 7608–7619
  • [20] Robert C. Reilly “Applications of the Hessian operator in a Riemannian manifold” In Indiana Univ. Math. J. 26.3, 1977, pp. 459–472
  • [21] Antonio Ros “Compact hypersurfaces with constant higher order mean curvatures” In Rev. Mat. Iberoamericana 3.3-4, 1987, pp. 447–453
  • [22] Julian Scheuer, Guofang Wang and Chao Xia “Alexandrov-Fenchel inequalities for convex hypersurfaces with free boundary in a ball” In J. Differential Geom. 120.2 Lehigh University, 2022, pp. 345–373
  • [23] Chunquan Tang “Mixed boundary value problems for quasilinear elliptic equations” Thesis (Ph.D.)–Iowa State University ProQuest LLC, Ann Arbor, MI, 2013, pp. 80
  • [24] Guofang Wang, Liangjun Weng and Chao Xia “Alexandrov–Fenchel inequalities for convex hypersurfaces in the half-space with capillary boundary” In Math. Ann. Springer, 2023, pp. 1–34
  • [25] Guofang Wang and Chao Xia “Uniqueness of stable capillary hypersurfaces in a ball” In Math. Ann. 374.3-4, 2019, pp. 1845–1882
  • [26] Liangjun Weng and Chao Xia “Alexandrov-Fenchel inequality for convex hypersurfaces with capillary boundary in a ball” In Trans. Amer. Math. Soc. 375.12, 2022, pp. 8851–8883
  • [27] Thomas Young “An essay on the cohesion of fluids” In Philos. Trans. Roy. Soc. A The Royal Society London, 1805, pp. 65–87