Skip to main content
Log in

A Heintze–Karcher-Type Inequality for Hypersurfaces with Capillary Boundary

  • Published:
The Journal of Geometric Analysis Aims and scope Submit manuscript

Abstract

In this paper, we establish a Heintze–Karcher-type inequality for hypersurfaces with capillary boundary of contact angle \(\theta \in \big (0,\frac{\pi }{2}\big ]\) in the half-space or a half ball, by using solution to a mixed boundary value problem in Reilly type formula. Consequently, we give a new proof of Alexandrov-type theorem for embedded capillary constant mean curvature hypersurfaces with contact angle \(\theta \in \big (0,\frac{\pi }{2}\big ]\) in the half-space or a half ball.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Quite recently, the conjecture in the half-space case has been solved by the authors together with Wang in [8], see also [9] for the extension to the anisotropic setting.

References

  1. Ainouz, A., Souam, R.: Stable capillary hypersurfaces in a half-space or a slab. Indiana Univ. Math. J. 65, 813–831 (2016). https://doi.org/10.1512/iumj.2016.65.5839

    Article  MathSciNet  MATH  Google Scholar 

  2. Aleksandrov, A.D.: Uniqueness theorems for surfaces in the large I, II. Transl. Ser. 2 Am. Math. Soc. 21, 341–354 (1962). https://doi.org/10.1090/trans2/021/09

    Article  MathSciNet  MATH  Google Scholar 

  3. Athanassenas, M.: A variational problem for constant mean curvature surfaces with free boundary. J. Reine Angew. Math. 377, 97–107 (1987). https://doi.org/10.1515/crll.1987.377.97

    Article  MathSciNet  MATH  Google Scholar 

  4. Choe, J., Park, S.-H.: Capillary surfaces in a convex cone. Math. Z. 267, 875–886 (2011). https://doi.org/10.1007/s00209-009-0651-3

    Article  MathSciNet  MATH  Google Scholar 

  5. Gilbarg, D., Trudinger, N.S.: Elliptic Partial Differential Equations of Second Order, p. 517. Springer, Berlin (2001)

    Book  MATH  Google Scholar 

  6. Guo, J., Xia, C.: A partially overdetermined problem in a half ball. Calc. Var. Partial Differ. Equ. 58, 15 (2019). https://doi.org/10.1007/s00526-019-1603-3

    Article  MathSciNet  MATH  Google Scholar 

  7. Heintze, E., Karcher, H.: A general comparison theorem with applications to volume estimates for submanifolds. Ann. Sci. Norm. Supér. 11, 451–470 (1978). https://doi.org/10.24033/asens.1354

  8. Jia, X., Wang, G., Xia, C., Zhang, X.: Alexandrov’s theorem for anisotropic capillary hypersurfaces in the half-space. Arch. Ration. Mech. Anal. (to appear). Preprint at http://arxiv.org/abs/2211.02913 (2022)

  9. Jia, X., Wang, G., Xia, C., Zhang, X.: Heintze–Karcher inequality and capillary hypersurfaces in a wedge. Preprint at http://arxiv.org/abs/2209.13839 (2022)

  10. Li, J., Xia, C.: An integral formula and its applications on sub-static manifolds. J. Differ. Geom. 113, 493–518 (2019). https://doi.org/10.4310/jdg/1573786972

    Article  MathSciNet  MATH  Google Scholar 

  11. Liebermann, G.M.: Mixed boundary value problems for elliptic and parabolic differential equations of second order. J. Math. Anal. Appl. 113, 422–440 (1986). https://doi.org/10.1016/0022-247X(86)90314-8

    Article  MathSciNet  Google Scholar 

  12. Liebermann, G.M.: Optimal Hölder regularity for mixed boundary value problems. J. Math. Anal. Appl. 143, 572–586 (1989). https://doi.org/10.1016/0022-247X(89)90061-9

    Article  MathSciNet  MATH  Google Scholar 

  13. López, R.: Capillary surfaces with free boundary in a wedge. Adv. Math. 262, 476–483 (2014). https://doi.org/10.1016/j.aim.2014.05.019

    Article  MathSciNet  MATH  Google Scholar 

  14. Maggi, F.: Sets of Finite Perimeter and Geometric Variational Problems. An Introduction to Geometric Measure Theory, vol. 135, p. 454. Cambridge University Press, Cambridge (2012)

    Book  MATH  Google Scholar 

  15. Pacella, F., Tralli, G.: Overdetermined problems and constant mean curvature surfaces in cones. Rev. Mater. Iberoam. 36, 841–867 (2020). https://doi.org/10.4171/rmi/1151

    Article  MathSciNet  MATH  Google Scholar 

  16. Qiu, G., Xia, C.: A generalization of Reilly’s formula and its applications to a new Heintze–Karcher type inequality. Int. Math. Res. Not. 2015, 7608–7619 (2015). https://doi.org/10.1093/imrn/rnu184

    Article  MathSciNet  MATH  Google Scholar 

  17. Reilly, R.C.: Applications of the Hessian operator in a Riemannian manifold. Indiana Univ. Math. J. 26, 459–472 (1977). https://doi.org/10.1512/iumj.1977.26.26036

    Article  MathSciNet  MATH  Google Scholar 

  18. Ros, A.: Compact hypersurfaces with constant higher order mean curvatures. Rev. Mater. Iberoam. 3, 447–453 (1987). https://doi.org/10.4171/RMI/58

    Article  MathSciNet  MATH  Google Scholar 

  19. Ros, A., Souam, R.: On stability of capillary surfaces in a ball. Pac. J. Math. 178, 345–361 (1997). https://doi.org/10.2140/pjm.1997.178.345

    Article  MathSciNet  MATH  Google Scholar 

  20. Serrin, J.: A symmetry problem in potential theory. Arch. Ration. Mech. Anal. 43, 304–318 (1971). https://doi.org/10.1007/BF00250468

    Article  MathSciNet  MATH  Google Scholar 

  21. Souam, R.: On stable capillary hypersurfaces with planar boundaries. Preprint at http://arxiv.org/abs/2111.01500 (2021)

  22. Tang, C.: Mixed boundary value problems for quasilinear elliptic equations. Thesis (Ph.D.)-Iowa State University. ProQuest LLC, Ann Arbor, pp. 80 (2013). http://gateway.proquest.com/openurl?url_ver=Z39.88-2004 &rft_val_fmt=info:ofi/fmt:kev:mtx:dissertation &res_dat=xri:pqm &rft_dat=xri:pqdiss:3566146

  23. Wang, G., Xia, C.: Uniqueness of stable capillary hypersurfaces in a ball. Math. Ann. 374, 1845–1882 (2019). https://doi.org/10.1007/s00208-019-01845-0

    Article  MathSciNet  MATH  Google Scholar 

  24. Wente, H.C.: The symmetry of sessile and pendent drops. Pac. J. Math. 88, 387–397 (1980). https://doi.org/10.2140/pjm.1980.88.387

    Article  MathSciNet  MATH  Google Scholar 

Download references

Funding

This work is supported by the NSFC (Grant Nos. 11871406, 12271449, 12126102).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chao Xia.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A: A Fredholm Alternative for the Mixed Boundary Value Problem

Appendix A: A Fredholm Alternative for the Mixed Boundary Value Problem

The purpose of the appendix is to present a detailed statement and proof of the Fredholm alternative for mixed boundary elliptic equation, which was brought up in [11] without proof.

For completeness, we present some existence and regularity results for mixed boundary value problems

$$\begin{aligned} {\left\{ \begin{array}{ll} {{\bar{\Delta }}} f=h \quad &{}\text {in }\Omega ,\\ f=0 \quad &{}\text {on }\Sigma ,\\ {{\bar{\nabla }}}_{{{\bar{N}}}} f+ f=g &{}\text {on }\textrm{int}(T), \end{array}\right. } \end{aligned}$$
(48)

which was proved by Lieberman [11, 12].

To facilitate the presentation, We recall the definition of weighted Hölder spaces. Set \(d_\Gamma (x)=\textrm{dist}(x,\Gamma )\), \(\Omega _\delta =\{x\in \Omega :d_\Gamma (x)>\delta \}\). For \(a\ge 0\), \(b\ge -a\), we define

$$\begin{aligned} {|f|}_a^{(b)}=\sup _{\delta >0}\delta ^{a+b}{|f|}_{a;\Omega _\delta }, \end{aligned}$$

where \(|f|_{a;\Omega _\delta }\) is the standard norm on \(\Omega _{\delta }\). We denote by \(H_a^{(b)}\) the set of all functions f on \(\Omega \) with finite norm \(|f|_a^{(b)}\).

Theorem A.1

There exists a solution \(f\in C^{2}( \Omega \cup \textrm{int} (T))\cap C^0({{\overline{\Omega }}})\) to (48) for all \(h\in C^{\alpha }(\Omega \cup \textrm{int} (T))\), \(g\in C^{1,\alpha }(\Omega \cup \textrm{int} (T))\).

Proof

See [11, Theorem 1]. \(\square \)

Theorem A.2

Assume \(\theta \in (0,\frac{\pi }{2})\). Let \(f\in C^{2}( \Omega \cup \textrm{int} (T))\cap C^{0}({{\overline{\Omega }}})\) be a solution to (48). Then for any \(\lambda \in \left( 1,\frac{\pi }{2\theta }\right) \) and any noninteger \(a>2\), we have

$$\begin{aligned} |f|_a^{(-\lambda )}\le C\bigg (|h|_{a-2}^{(2-\lambda )}+|g|_{a-1}^{(1-\lambda )}+|f|_0\bigg ). \end{aligned}$$
(49)

Proof

See [12, Theorem 4]. Here we only explain the admissible range \(\left( 1,\frac{\pi }{2\theta }\right) \) for \(\lambda \), which is not explicitly expressed in [12, Theorem 4].

Locally near every \(x_0\in \Gamma \), up to a transformation, there exists a Cartesian coordinate \((x^1,\ldots ,x^{n+1})\), centered at \(x_0\), such that the corresponding cylindrical coordinates \((r,\eta ,x')\) is given by

$$\begin{aligned} x_1=r\cos \eta ,\quad x_2=r\sin \eta ,\quad x'=\big (x^3,\ldots ,x^{n+1}\big ), \end{aligned}$$

with \(0\le \eta \le \theta \). This follows from the definition of the wedge condition by Lieberman in [11], which is clearly satisfied by domains that we consider.

The key of the proof in [12, Theorem 4] is to find a Miller-type barrier function \(\psi =r^\lambda \varphi (\eta )\) satisfying

$$\begin{aligned} {\left\{ \begin{array}{ll} {{\bar{\Delta }}} \psi =r^{\lambda -2}(\lambda ^2\varphi +\varphi _{\eta \eta })\le -c_1r^{\lambda -2}\quad &{}\text {for }\eta \in (0,\theta ),\\ {{\bar{\nabla }}}_{{{\bar{N}}}} \psi =-r^{\lambda -1}\varphi _{\eta }\ge c_1 r^{\lambda -1} &{}\text {for } \eta =0,\\ c_1\le \varphi \le 1 &{}\text {for } \eta \in [0,\theta ], \end{array}\right. } \end{aligned}$$
(50)

with a positive constant \(c_1\). The Miller-type barrier \(\varphi (\eta )\) is constructed from a perturbation of \({\tilde{\varphi }}(\eta )=\cos (\lambda \eta )\), which has a positive lower bound in \([0, \theta ]\) and satisfies (50) only if \(\theta \in (0,\frac{\pi }{2\lambda })\). Hence to ensure \(\lambda >1\), one needs to restrict \(\theta \in (0,\frac{\pi }{2})\). \(\square \)

Lemma A.1

If \(f\in C^{2}( \Omega \cup \textrm{int} (T))\cap C^{0}({{\overline{\Omega }}})\) is a solution to (48). Then there exists a positive constant C such that

$$\begin{aligned} |f|_0\le C(|h|_0+|g|_0). \end{aligned}$$
(51)

Proof

Set \(C_1=|f|_0+|g|_0\). For the case \(B=\overline{{\mathbb {R}}^{n+1}_+}\), assume without loss of generality that \(\Omega \subset \left\{ 0<x_{n+1}<R\right\} \) for some R large, consider the functions \(v_1(x)=C_1(e^{x_{n+1}}-e^R)\). A direct computation then yields,

$$\begin{aligned} {\left\{ \begin{array}{ll} {{\bar{\Delta }}} v_1\ge C_1 \quad &{}\text {in }\Omega ,\\ v_1\le 0 \quad &{}\text {on }\Sigma ,\\ {{\bar{\nabla }}}_{{{\bar{N}}}} v_1+v_1\le -C_1 &{}\text {on } \textrm{int}(T). \end{array}\right. } \end{aligned}$$
(52)

Applying the maximum principle (see for example [22, Lemma 4.1]) to \(f+v_1\) and \(-f+v_1\), we obtain (51).

For the case \(B=\overline{{\mathbb {B}}^{n+1}}\), we choose \(v_2=C_1(|x|^2-4)\) to replace \(v_1\), and the process follows similarly. \(\square \)

We can now derive a Fredholm alternative for the mixed boundary value problem.

Theorem A.3

Assume \(\theta \in \big (0,\frac{\pi }{2}\big )\), let a be a noninteger greater than 2, \(\lambda \) be defined as in Theorem A.2. Then either (a) the homogeneous problem

$$\begin{aligned} {\left\{ \begin{array}{ll} {{\bar{\Delta }}} f=0 \quad &{}\text {in }\Omega ,\\ f=0 \quad &{}\text {on }\Sigma ,\\ {{\bar{\nabla }}}_{{{\bar{N}}}} f-\gamma f=0 &{}\text {on }\textrm{int}(T), \end{array}\right. } \end{aligned}$$
(53)

has nontrivial solutions; or (b) the homogeneous problem has only the trivial solution, in which case, for all \({h}\in H_{a-2}^{(2-\lambda )}\), \(g\in H_{a-1}^{1-\lambda }\) the inhomogeneous problem

$$\begin{aligned} {\left\{ \begin{array}{ll} {{\bar{\Delta }}} f=h \quad &{}\text {in }\Omega ,\\ f=0 \quad &{}\text {on }\Sigma ,\\ {{\bar{\nabla }}}_{{{\bar{N}}}} f-\gamma f=g &{}\text {on }\textrm{int}(T), \end{array}\right. } \end{aligned}$$
(54)

has a unique solution \(f\in H_a^{(-\lambda )}\).

Remark A.1

By the monotonicity of norm we have \(|f|_\lambda ^{(-\lambda )}\le |f|_a^{(-\lambda )}\). Therefore if there is a solution \(f\in H_{a}^{(-\lambda )}\) to problem (54), then the solution f is also in \(C^{1,\alpha }({{\overline{\Omega }}})\) with \(\alpha =\lambda -1\).

Proof of Theorem A.3

We follow the classical proof of the Fredholm alternative for the Dirichlet problem but with a careful choice of function space. Set \({\mathcal {A}} =\{u\in H_a^{(-\lambda )}: f=0\text { on }\Sigma \}\), \({\mathcal {B}}= H_{a-2}^{(2-\lambda )}\times H_{a-1}^{(1-\lambda )}\). By Theorem A.1, Theorem A.2 and (51), there exists a solution \(u\in {\mathcal {A}}\) to problem (48) for all \((h,g)\in {\mathcal {B}}\). Uniqueness of the solution to problem (48) follows from the maximum principle immediately. Let \(Q:{\mathcal {B}}\rightarrow {\mathcal {A}}\) be an operator such that Q(hg) is the unique solution of problem (48). One can readily see that the operator Q is well-defined and bijective. Notice that the inhomogeneous problem (54) is equivalent to the following equation

$$\begin{aligned} f-Q(0,(1+\gamma )f)=Q(h,g). \end{aligned}$$
(55)

Thus \(f\in {\mathcal {A}}\) is a solution to (55) if and only if it solves (54). Let P be an operator from \(H_{a-1}^{(1-\lambda )}\) to itself such that \(Pu=Q(0,(1+\gamma )u)\) for all \(u\in H_{a-1}^{(1-\lambda )}\). Then the equation (55) reads as

$$\begin{aligned} f-Pf=v, \end{aligned}$$
(56)

where \(v=Q(h,g)\).

To apply the classical Fredholm alternative (see e.g., [5, Theorem 5.11]) for (56), we need to verify that P is a compact operator. To this end, let \(\{g_k\}\) be a bounded sequence in \(H_{a-1}^{(1-\lambda )}\). By Theorem A.1 and (51), there exists \(f_k\in {\mathcal {A}}\) such that \(Pg_k=f_k\), and we have

$$\begin{aligned} |f_k|_a^{(-\lambda )}\le C\left( |g_k|_{a-1}^{(1-\lambda )}+|f_k|_0\right) . \end{aligned}$$
(57)

Since \(\left\{ |f_k|_0\right\} _k\) is bounded, we have \(\left\{ f_k\right\} _k\) is also bounded in \(H_{a}^{(-\lambda )}\). By virtue of the Ascoli-Arzela theorem, up to a subsequence, there exists \({{\tilde{f}}}\in H_{a-1}^{(1-\lambda )}\), such that \(f_{k_j}\) converges to \({{\tilde{f}}}\) in \(H_{a-1}^{(1-\lambda )}\). Thus P is compact and the classical Fredholm alternative applies: (56) has a unique solution \(f\in H_{a-1}^{(1-\lambda )}\), provided that the homogenous equation \(f-Pf=0\) has only the trivial solution \(f=0\). Since Q maps \({\mathcal {B}}\) onto \({\mathcal {A}}\), any solution \(f\in H_{a-1}^{(1-\lambda )}\) of (56) also belongs to \({\mathcal {A}}\). This completes the proof. \(\square \)

Remark A.2

We remark that, in the case \(\theta =\frac{\pi }{2}\), we can use boundary reflection to get better regularity, say global \(W^{2, p}\)-estimate for any p, see for example [6, Proposition 3.5].

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jia, X., Xia, C. & Zhang, X. A Heintze–Karcher-Type Inequality for Hypersurfaces with Capillary Boundary. J Geom Anal 33, 177 (2023). https://doi.org/10.1007/s12220-023-01230-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12220-023-01230-z

Keywords

Mathematics Subject Classification

Navigation